Fact-checked by Grok 2 weeks ago

Proton (subatomic particle)

The proton is a subatomic particle with a positive equal to the of +1.602176634 × 10^{-19} coulombs and a rest mass of 1.67262192595 × 10^{-27} kilograms, making it the lightest known and a fundamental constituent of all nuclei. Protons, along with neutrons, form the dense core of every , where they are bound by the ; the number of protons in a , known as the , uniquely determines the and governs its chemical properties. In the of , protons are composite particles classified as hadrons, specifically baryons composed of three valence quarks—two up quarks and one —held together by gluons that mediate the strong interaction, with the proton's positive charge arising from the +2/3 charges of the two up quarks and the -1/3 charge of the , for a net charge of +1. The existence of the proton as a distinct particle was established through experiments conducted by between 1917 and 1919, in which bombardment of gas produced nuclei identifiable as positively charged particles from atomic cores. Protons also exhibit an intrinsic or of ½ ħ, a property that contributes to the overall structure of matter and has been the subject of ongoing research into the "proton spin crisis," where the distribution of among quarks, gluons, and orbital motion remains an active area of investigation.

Fundamental Properties

Composition and Structure

The proton is a , a type of composed of three valence quarks within the of : specifically, two up quarks (u) and one (d), denoted by the quantum numbers uud. These valence quarks carry the proton's of +1 and determine its fundamental quantum properties, such as flavor and charge, while being confined within the proton due to the strong interaction. The binding of these quarks occurs through the exchange of gluons, the gauge bosons of (QCD), the theory describing the . In QCD, quarks possess a (red, green, or blue), and gluons mediate the , ensuring that the overall proton is a color-neutral , as isolated colored particles cannot exist freely due to . This dynamic interaction generates the proton's structure, with gluons carrying a significant portion of the proton's —approximately half at typical scales probed in experiments. Beyond the quarks, the proton's internal includes a "" of virtual quark-antiquark pairs and gluons, which contribute to its parton distribution functions as revealed by (DIS) experiments. In DIS, high-energy electrons or muons scatter off the proton, probing its constituents at short distances; the resulting structure functions, such as F₂(x, Q²), show that sea quarks (including flavors like strange and ) and gluons make up the remaining momentum fraction not accounted for by valence quarks, with their distributions evolving with the momentum transfer Q² according to QCD predictions. These virtual particles arise from quantum fluctuations, complicating the proton's composition but essential for understanding its behavior in high-energy collisions. The proton's in QCD incorporates these elements, describing the for finding quarks of specific and colors in particular configurations. For the ground-state proton, the spatial wave function is symmetric, combined with and flavor parts to yield a total antisymmetric color wave function that is a under SU(3)color, ensuring confinement; in the simple approximation, it is often expressed as a combination like (1/√18) ∑ ε{ijk} (u_i ↑ u_j ↓ d_k ↑ - u_i ↑ u_j ↑ d_k ↓ + permutations), where indices denote color and arrows , though full QCD calculations involve effects. The proton's mass arises predominantly from this QCD rather than the bare masses of the quarks.

Mass, Charge, and Spin

The rest of the proton is 938.27208816(29) MeV/c². Approximately 99% of this mass originates from the binding energy due to (QCD) interactions among its quarks and gluons, with only about 1% attributable to the rest masses of the constituent quarks. The proton carries an of +1 e, equivalent to +1.602176634 × 10⁻¹⁹ C, which defines the fundamental unit of positive charge in the atomic scale and balances the negative charge of the in neutral atoms. The proton possesses an intrinsic spin angular momentum of ½ ħ, classifying it as a fermion and dictating its behavior under the Pauli exclusion principle in quantum systems. This spin gives rise to a magnetic dipole moment of +2.79284734(3) nuclear magnetons, which exceeds the Dirac value of 1 nuclear magneton expected for a point-like spin-½ particle due to its composite structure. The magnetic moment is expressed by the formula \mu_p = g_p \mu_N \frac{S}{\hbar}, where g_p \approx 5.585694689 is the proton's g-factor, \mu_N = e \hbar / (2 m_p) is the nuclear magneton with proton mass m_p, and S is the spin operator.

Charge Radius and Size

The proton's finite size is quantified by its charge radius, which reflects the spatial distribution of its electric charge rather than treating it as a point particle. The root-mean-square (RMS) charge radius is given by \sqrt{\langle r^2 \rangle} = \left( \frac{\int r^2 \rho(\mathbf{r}) \, d^3\mathbf{r}}{\int \rho(\mathbf{r}) \, d^3\mathbf{r}} \right)^{1/2}, where \rho(\mathbf{r}) denotes the charge density function. Measurements of this radius have converged on a value of approximately 0.841 fm, obtained through complementary techniques such as elastic electron-proton scattering at facilities like Jefferson Lab and spectroscopy of muonic hydrogen at the Paul Scherrer Institute. These methods probe the proton's charge distribution at different length scales, with electron scattering providing form factor data and muonic spectroscopy exploiting the muon's tighter orbit for enhanced sensitivity. A notable challenge in these determinations was the "proton radius puzzle," which emerged around 2010 when muonic hydrogen spectroscopy yielded a smaller radius of about 0.841 fm compared to roughly 0.877 fm from earlier electron-based measurements. This 4–7σ discrepancy prompted extensive scrutiny of systematic effects, theoretical corrections, and new experiments. By the early 2020s, refined electronic spectroscopy and scattering data had reduced the tension, confirming the muonic value of ~0.84 fm as the consensus, with ongoing efforts aiming for sub-1% precision. Lattice (QCD) simulations offer insights into the proton's internal structure, revealing a pressure profile driven by quark-gluon dynamics. The radial pressure distribution peaks at the proton's core with an immense value of ~$3 \times 10^{35} —far exceeding pressures in stars—before transitioning to repulsive forces near the center and tensile (negative) forces at the periphery. This profile, computed from the proton's energy-momentum tensor, underscores the intense confinement within the ~1 scale.

Stability and Lifetime

Experimental Limits

The experimental searches for proton decay utilize large underground detectors to isolate rare events from overwhelming backgrounds, primarily cosmic rays and atmospheric neutrinos. (SK), a 50-kiloton Cherenkov detector situated 1,000 meters underground in the Kamioka mine in , has provided the world's most stringent limits since its inception in 1996. The technique relies on detecting emitted by charged particles from potential decay products, reconstructing event topologies to identify signatures like a and , while vetoing muons to suppress backgrounds. With cumulative exposures of up to 0.45 megaton-years in recent analyses (data through 2018) and ongoing accumulation exceeding 0.42 megaton-years as of 2025, no candidates have been observed. Key searches target supersymmetry-favored modes predicted by grand unified theories. For the decay p \to e^+ \pi^0, SK analysis of data through 2018 yields a lower on the partial lifetime of \tau / B > 2.4 \times 10^{34} years at 90% level, improving prior bounds by incorporating an enlarged fiducial volume of 27.2 kilotons. Similarly, for p \to \mu^+ K^0, the same exposure sets \tau / B > 3.6 \times 10^{33} years at 90% CL, leveraging advanced identification via delayed Cherenkov rings from K^0_S \to \pi^+ \pi^- decays. These results stem from meticulous background modeling, achieving rejection rates below $10^{-5} for neutrino-induced events mimicking decays. The next-generation (Hyper-K), currently under construction with a 260-kiloton fiducial volume, will enhance sensitivity by an through increased target mass and upgraded photosensors. As of 2025, excavation of the main cavern has been completed. Projected limits include \tau / B > 10^{35} years for p \to e^+ \pi^0 and \tau / B > 3 \times 10^{34} years for p \to \nu \bar{K}^0 after 10 years of operation starting around 2027, enabling deeper probes into violation scales. An observation of proton decay would confirm grand unified theories by evidencing non-conservation, unifying the strong, weak, and electromagnetic forces at high energies. The proton's stability reflects conservation in the , motivating these null results to constrain physics beyond it.

Theoretical Implications

In the of , the proton's stability arises from the conservation of , a global U(1)_B that assigns B=1 to the proton and prevents its decay into non-baryonic particles such as leptons or mesons. This conservation is accidental, emerging because no renormalizable terms in the violate , despite the in the theory at high energies. The gauge structure of (QCD), with its SU(3)_c color , and the electroweak sector, governed by SU(2)_L × U(1)_Y, enforces this stability by prohibiting baryon-number-violating interactions at the renormalizable level, ensuring the proton—as the lightest —has no kinematically allowed decay channels that conserve charge, energy, and other quantum numbers. Grand unified theories (GUTs) extend the by unifying the strong, weak, and electromagnetic forces under a single gauge group, such as SU(5) proposed by Georgi and Glashow, where is no longer conserved. In the minimal SU(5) GUT, proceeds through the exchange of gauge bosons X and Y, which mediate transitions between quarks and leptons, enabling processes like p → e⁺ + π⁰ with ΔB = -1. The decay width for such processes is dimensionally estimated as \Gamma \approx \frac{\alpha^2 m_p^5}{m_X^4}, where α is the unified coupling constant (α ≈ 1/40), m_p is the proton mass, and m_X is the mass of the X/Y bosons, typically at the GUT scale. The corresponding proton lifetime is then τ = 1/Γ ≈ (m_X / 10^{15} \text{ GeV})^4 / \alpha^2 years, linking the decay rate directly to the unification scale m_X and providing a testable prediction that constrains GUT models. Experimental lower limits on the proton lifetime thus impose bounds on m_X, often requiring it to exceed 10^{15} GeV in viable GUTs.

Historical Development

Early Discoveries

The early investigations into the proton began with J.J. Thomson's studies of and positive ions in the late and early 1900s. Thomson's experiments with discharge tubes revealed streams of positively charged particles, known as canal rays or positive rays, which he analyzed using magnetic and electric fields to determine their mass-to-charge ratios. These observations demonstrated the existence of positive ions, including those from , providing the first experimental evidence for positively charged subatomic constituents that would later be identified as protons. Building on this foundation, Ernest Rutherford's 1911 gold foil experiment marked a pivotal advancement by revealing the . In this work, conducted with and , alpha particles were directed at a thin foil, and their scattering patterns were observed; the unexpected large-angle deflections indicated a tiny, dense, positively charged core at the atom's center, overturning Thomson's and establishing the nuclear model of the atom. This discovery implied that the positive charge and most of the mass resided in the , setting the stage for identifying its fundamental components. Rutherford's subsequent experiments from 1917 to further characterized the proton through alpha-particle bombardment of light elements. During , Rutherford resumed scattering studies at the , observing long-range recoils of particles when alpha particles interacted with and other light atoms like , , sodium, aluminum, and . In a key publication, he detailed how alpha particles penetrated nuclei, ejecting nuclei with high energy, confirming that these nuclei were fundamental building blocks present in other atomic nuclei. These findings, published in four papers including "An anomalous effect in nitrogen," demonstrated the ejection of protons via interactions, solidifying their role as elementary particles. In 1920, Rutherford formally named the hydrogen nucleus the "proton," derived from the Greek word "protos" meaning "first," recognizing it as the primary positive constituent of atomic nuclei. This nomenclature appeared in his Bakerian Lecture, emphasizing the proton's foundational importance in atomic structure.

Quark Model and Modern Understanding

The quark model emerged in 1964 as a theoretical framework to describe the structure of hadrons, including the proton. Independently proposed by Murray Gell-Mann and George Zweig, it posited that baryons like the proton consist of three fundamental constituents called quarks, with the proton specifically composed of two up quarks and one down quark (uud) to account for its charge and other quantum numbers. This model resolved puzzles in the spectroscopy of hadrons observed in the early 1960s, such as the pattern of masses and decays in the SU(3) flavor symmetry group, by introducing fractionally charged, spin-1/2 particles bound by strong interactions. Experimental confirmation of point-like quarks within the proton came from deep inelastic scattering experiments at the Stanford Linear Accelerator Center (SLAC) starting in 1968, led by Jerome Friedman, Henry Kendall, and Richard Taylor. These experiments involved high-energy electrons scattered off protons, revealing inelastic events consistent with interactions with small, point-like constituents carrying a fraction of the proton's momentum, as predicted by the parton model inspired by the quark hypothesis. The scaling behavior observed in the cross-sections supported the existence of these quarks as quasi-free partons at short distances. The theoretical foundation for the strong interactions binding quarks in the proton was solidified in the 1970s with the development of (QCD), formulated as a non-Abelian based on the SU(3) color group. and , along with independently David Politzer, demonstrated in 1973 that QCD exhibits , wherein the strong coupling constant decreases at high energies (short distances), enabling perturbative calculations for processes like . This property explained why quarks behave as nearly free partons at high energies while remaining confined at low energies, unifying the with the of the strong force. From the 1980s onward, has provided a computational approach to refine the proton's internal structure, discretizing into a lattice to simulate and dynamics via methods. Early lattice calculations focused on masses and matrix elements, yielding insights into the proton's and parton distributions that complement perturbative QCD. These simulations have progressively improved, incorporating finer lattices and lighter masses to accurately predict properties like the proton's and flavor-singlet contributions to its spin.

Natural Occurrence

In Atomic Nuclei

In atomic nuclei, protons serve as one of the two primary types of nucleons, alongside neutrons, forming the dense core of every atom. The number of protons in a nucleus defines the atomic number Z, which determines the element's identity and equals the number of electrons in a neutral atom. Protons are bound within the nucleus by the strong nuclear force, a fundamental interaction that acts over extremely short ranges of about 1 femtometer and is powerful enough to overcome the electromagnetic Coulomb repulsion between positively charged protons. This binding is essential for nuclear stability, as the Coulomb force would otherwise cause the protons to disperse. For instance, in the deuterium nucleus (hydrogen-2), one proton and one neutron are held together by this force, forming the simplest bound nuclear system beyond a single proton. Isotopes of an are variants that share the same number of protons but differ in count, with the proton number fixing the . The most common of , hydrogen-1 (protium), consists solely of a single proton in its , making it the only without neutrons and exemplifying a "bare" proton in terms. Protons contribute to overall stability through models like the liquid drop model, which approximates the as a charged liquid drop and calculates as a balance of attractive forces and repulsive effects among protons. In this semi-empirical framework, the binding energy B for a with Z protons and A includes a Coulomb term proportional to Z(Z-1)/A^{1/3}, highlighting how additional protons increase instability for larger nuclei unless balanced by neutrons. This model successfully predicts barriers and energetics for heavy elements.

In the Universe and Cosmic Rays

Protons are the most abundant baryons in the , comprising approximately 75% of the total baryonic matter by mass in the form of nuclei, a composition established during in the first few minutes after the . This primordial abundance arises from the rapid fusion of protons into light elements like , with the remaining ~25% being helium and trace amounts of , , and lithium-7. Observations of radiation and light element ratios confirm this distribution, providing key evidence for the standard cosmological model. In stellar interiors, protons play a central role in the proton-proton (pp) chain, the dominant mechanism in low-mass stars like , which converts into and releases energy essential for stellar stability. The pp chain proceeds through three main branches—ppI, ppII, and ppIII—with the ppI branch accounting for about 86% of reactions in , while ppII and ppIII contribute smaller fractions depending on temperature and abundance. The initial step of the chain involves the between two protons: p + p \rightarrow {}^2\mathrm{H} + e^+ + \nu_e This positron-electron annihilation and neutrino emission follow, enabling subsequent steps that ultimately produce helium-4 from four protons, powering the Sun's luminosity. Beyond stellar fusion, protons dominate primary cosmic rays, making up over 90% of their composition as high-energy particles originating from astrophysical accelerators. These protons are primarily accelerated in supernova remnants within our galaxy for lower energies and in active galactic nuclei for extragalactic contributions, achieving maximum energies around $10^{20} eV—the highest observed for any particle. This proton flux, detected at Earth after interactions with interstellar medium, provides insights into extreme astrophysical environments and particle acceleration mechanisms.

Interactions

With Ordinary Matter

When free protons traverse ordinary matter, their primary interaction is through collisions with electrons, leading to where electrons are stripped from atoms, producing ion pairs and . This process dominates energy loss for protons with energies above approximately 1 keV, as the proton's is much larger than the electron's, resulting in small momentum transfer per collision but cumulative along the path. In denser media or gases, repeated ionizations can form molecular ions such as H₂⁺ via charge exchange with atoms, or at sufficiently high proton fluxes, generate partially ionized states. These interactions are purely electromagnetic at low energies, with negligible nuclear contributions below several MeV. For low-energy protons (typically below 10 MeV), off atomic nuclei via the interaction governs angular deflection, described by the formula. The differential cross-section for such is \frac{d\sigma}{d\Omega} = \left( \frac{Z e^2}{8\pi \epsilon_0 E} \right)^2 \frac{1}{\sin^4 \left( \frac{\theta}{2} \right)}, where Z is the of the target nucleus, e is the , \epsilon_0 is the , E is the proton's , and \theta is the scattering angle in the lab frame (approximating the center-of-mass frame for heavy targets). This formula predicts a strong forward-peaking distribution, with large-angle scatters rare due to the $1/r potential. Multiple small-angle scatters accumulate to broaden the proton's trajectory, but single large deflections remain possible. The of protons in matter is determined by their energy loss rate, primarily via the Bethe-Bloch formula, which quantifies the mean loss per unit path length: -\frac{dE}{dx} = \frac{4\pi z^2 e^4 N_A Z \rho}{(4\pi \epsilon_0)^2 m_e c^2 \beta^2 A} \left[ \ln \left( \frac{2 m_e c^2 \beta^2 \gamma^2}{I} \right) - \beta^2 \right], where z = 1 for protons, N_A is Avogadro's number, \rho and A are the target's and atomic mass, m_e is the , \beta = v/c, \gamma = 1/\sqrt{1 - \beta^2}, and I is the mean excitation energy (relativistic approximation, ignoring higher-order terms). This results in a finite range, with protons slowing gradually until stopping; for instance, a 1 MeV proton penetrates approximately 0.15 mm in before halting, depositing most energy near the end via a . These and processes are harnessed in particle detectors, particularly ionization chambers, where protons passing through a gas-filled volume (e.g., or air) produce measurable pairs proportional to their loss. The collected charge yields information on proton , , and position, with applications in beam monitoring and ; for example, parallel-plate chambers in proton accelerators detect beam profiles nondestructively. Such detectors operate reliably up to high fluxes, with corrections applied for recombination effects at elevated dose rates.

Electromagnetic and Nuclear Forces

The strong force that binds the quarks within a proton is mediated by the exchange of gluons, as described by (QCD), the theory of the strong interaction. This acts between color-charged quarks and gluons, exhibiting at short distances but confinement at larger scales, restricting its effective range to approximately 1 fm, the typical size of the proton. As a result, quarks cannot be isolated, and the proton remains a color-neutral . Protons carry a positive electric charge equal to the e, leading to Coulomb repulsion between two protons governed by the electromagnetic , with potential energy V(r) = \frac{e^2}{4\pi\epsilon_0 r}. In the context of atomic nuclei, this long-range repulsion is counteracted and effectively screened by the short-range attractive , primarily mediated by the exchange of virtual pions as proposed in Yukawa's meson-exchange theory. The resulting interaction follows a form V(r) \propto \frac{e^{-m_\pi r / \hbar c}}{r}, where m_\pi sets the range to about 1.4 , allowing stable multi-proton nuclei despite the electrostatic barrier. The weak force also influences proton-related processes, notably in , where a in a transforms into an , converting the neutron to a proton while emitting an and an electron antineutrino: n \to p + e^- + \bar{\nu}_e. This charged-current interaction, mediated by W bosons, violates parity and enables flavor-changing transitions essential for stellar nucleosynthesis and radioactive decay chains. Among these, the forces exhibit a clear hierarchy at nuclear scales: the strong force is approximately 100 times stronger than the electromagnetic force within distances of about 1 fm, while the weak force is much feebler, with coupling strength around $10^{-6} relative to the electromagnetic. This disparity ensures the strong force dominates quark binding and nuclear cohesion, with electromagnetic effects becoming prominent only at larger separations or in highly charged systems.

Role in Chemistry

As Hydrogen Ion

The proton serves as the nucleus of the , consisting of a single proton with no neutrons in its most common , protium, and thereby defining hydrogen's as 1 in the periodic table. This fundamental structure positions the proton as the central particle in hydrogen's chemical behavior, where it can be ionized to form the , a key species in acid-base chemistry. In aqueous solutions, the bare H⁺ proton does not exist freely but rapidly associates with water molecules to form the hydronium ion, H₃O⁺, through of H₂O. The acidity of such solutions is quantified by the scale, defined as pH = -log₁₀[H⁺], where [H⁺] represents the effective concentration of protons, influencing a wide range of chemical equilibria and reactions. According to the Brønsted-Lowry theory of acids and bases, introduced in , acids are substances that donate protons (H⁺), while bases accept them, emphasizing proton transfer as the core of acid-base reactions rather than electron pair interactions. For instance, dissociates in as HCl → H⁺ + Cl⁻, with the proton donated to H₂O to yield H₃O⁺, exemplifying a strong acid's complete proton transfer. Proton transfer reactions often exhibit rapid , with rate constants approaching the diffusion limit in , but in enzymatic contexts, these processes can involve quantum mechanical proton tunneling, where the proton passes through energy barriers rather than over them, enhancing reaction efficiency at physiological temperatures. This tunneling is evidenced by large kinetic isotope effects, such as substitution reducing rates by factors exceeding classical predictions, as observed in enzymes like and . Such quantum effects underscore the proton's role in facilitating ultrafast proton-coupled transfers critical for biological .

In Nuclear Magnetic Resonance

The proton's nuclear spin of I = \frac{1}{2} makes it particularly suitable for (NMR) , as this results in two distinct energy levels in an applied , enabling the absorption and emission of radiofrequency energy to produce observable signals. In ¹H NMR experiments, these signals occur at Larmor frequencies typically ranging from 400 to 900 MHz, corresponding to the strengths of standard superconducting magnets used in spectrometers, such as 9.4 T fields yielding about 400 MHz or 18.8 T fields approaching 800 MHz. This frequency range allows for high-resolution spectra that reveal detailed information about the proton's environment in molecules. A key feature in ¹H NMR is the chemical shift, denoted as \delta and expressed in parts per million (ppm), which reflects the influence of the local magnetic field on the proton due to surrounding electrons and molecular structure. Protons in different chemical environments experience shielding or deshielding effects, causing shifts typically between 0 and 12 ppm; for instance, methyl protons in alkanes appear around 0.9 ppm, while aldehydic protons resonate near 9-10 ppm. The standard reference compound is tetramethylsilane (TMS), assigned a chemical shift of 0 ppm, providing a consistent baseline for comparisons across samples. These shifts are independent of the strength, making \delta a robust measure for identifying functional groups. Spin-spin coupling, quantified by the scalar coupling constant J in hertz (Hz), arises from through-bond interactions between magnetically non-equivalent protons, typically over 1 to 3 bonds, and manifests as signal splitting in the spectrum. The multiplicity of peaks follows the n + 1 rule, where n is the number of equivalent neighboring protons; for example, a proton coupled to two equivalent protons appears as a triplet with J values around 7 Hz in ethyl groups. This coupling provides critical insights into proton-proton connectivity and , aiding in the elucidation of molecular frameworks. ¹H NMR is a cornerstone in for determining the structure of small molecules and complex natural products, where integration of peak areas quantifies proton counts, and the combination of chemical shifts and coupling patterns confirms connectivity without destroying the sample. Beyond the lab, the principles of proton NMR form the basis of (MRI), which detects signals from protons in and to generate high-contrast images of soft tissues , revolutionizing diagnostic since its development in the 1970s.

Applications and Biological Effects

Proton Therapy

Proton therapy, also known as proton beam therapy, is a form of external beam radiation treatment that utilizes accelerated protons to target cancerous tumors with high precision. Unlike conventional radiotherapy, where energy deposition decreases exponentially throughout the treatment volume, protons exhibit a characteristic energy deposition profile known as the . This peak occurs at a well-defined depth corresponding to the proton's range in tissue, beyond which the dose falls off sharply to nearly zero, minimizing exposure to healthy tissues distal to the tumor. The concept of using protons for was first proposed by physicist in 1946, recognizing their potential for precise dose delivery. To achieve therapeutic penetration depths of several centimeters to tens of centimeters in human , protons are accelerated to energies typically ranging from 70 to 250 MeV using particle accelerators such as cyclotrons or synchrotrons. For superficial tumors, lower energies suffice, while deeper-seated malignancies require higher energies to position the accurately within the target. To cover tumors with varying thicknesses, the proton beam undergoes range modulation, often via rotating ridge filters or adjustable energy degraders, which superimpose multiple pristine s of differing depths. This creates a spread-out Bragg peak (SOBP) that provides a uniform dose plateau across the tumor volume while maintaining the sharp distal falloff. The first clinical applications of emerged in the mid-1950s at research facilities like , where early prototypes demonstrated feasibility for treating ocular and pituitary tumors. By the , advancements in accelerator technology and treatment planning led to widespread adoption, with 108 operational proton therapy centers worldwide as of October 2025, predominantly equipped with compact superconducting cyclotrons for efficient beam production. Dosimetrically, the (LET) at the reaches approximately 20 keV/μm, higher than the ~0.2 keV/μm typical of X-rays, enhancing in the tumor while the overall dose profile reduces integral dose to surrounding organs, potentially lowering long-term side effects such as secondary cancers. Clinical evidence supports improved outcomes in pediatric cancers and skull base tumors due to this sparing effect.

Human Exposure and Radiation

Humans encounter proton radiation primarily through galactic cosmic rays (GCR) and solar particle events (), which pose significant risks during space travel due to the ionizing nature of high-energy protons. These exposures can lead to (), characterized by symptoms such as , , and potential fatality, particularly in unshielded or extravehicular activities. A historical example is the August 1972 , one of the most intense on record, which could have resulted in severe effects or for astronauts on lunar missions if they had coincided with the event. The biological impact of proton radiation is quantified using the in sieverts (), which incorporates the (RBE) to reflect tissue damage compared to reference like gamma rays. For protons, the RBE is approximately 1.1 in biological tissues, meaning proton doses are scaled by this factor to estimate equivalent effects. This adjustment helps assess risks from both acute high-dose events and chronic low-level exposures in space environments. Occupational exposure limits for radiation workers, including those handling proton sources in accelerators or medical facilities, are set by the (ICRP) at an effective dose of 20 mSv per year averaged over five consecutive years, with no single year exceeding 50 mSv to prevent effects like cancer. These limits apply broadly to , including protons, and are enforced to minimize long-term health risks. To mitigate proton components in GCR during deep-space missions, shielding often incorporates , a hydrogen-rich polymer that effectively attenuates high-energy protons and secondary neutrons through and fragmentation. studies demonstrate that polyethylene outperforms traditional aluminum shields by reducing dose equivalents by up to 30-50% for typical GCR proton fluxes, enhancing crew safety without excessive mass penalties.

Antiparticle

Antiproton Properties

The is the antiparticle of the proton, composed of two anti-up quarks and one anti-down quark (ūūd). It carries an of −1 e, a rest mass of 938.272 MeV/c² identical to that of the proton, and has 1/2. The was discovered in 1955 at the University of California's Berkeley Bevatron accelerator by a team led by and , along with Clyde Wiegand and Thomas Ypsilantis, through the identification of negatively charged particles with the proton's mass produced in proton-nucleus collisions. For this achievement, Chamberlain and Segrè shared the 1959 . The antiproton's magnetic moment has been measured as −2.79284734(42) μ_N, the negative counterpart to the proton's value, providing stringent confirmation of CPT symmetry with a fractional precision of 1.5 parts per billion. Like the proton, the antiproton is stable, with no observed decay; its mean lifetime is inferred from CPT symmetry to exceed 10^{34} years, the same lower limit as for the proton, although direct measurements from antiproton storage experiments yield >10 years and cosmic ray flux analyses provide limits around 10^7 years, though it annihilates upon contact with ordinary matter.

Production and Annihilation

Antiprotons are primarily produced in high-energy particle accelerators through the collision of protons with a fixed target, generating a cascade of secondary particles that includes antiprotons among pions, kaons, and other hadrons. At , protons accelerated to 26 GeV/c in the are directed onto an target, a proton-rich metal chosen for its high density and resistance to damage under intense bombardment. This process relies on mechanisms where the of the incoming proton exceeds the 1.876 GeV needed to create a proton-antiproton pair, with antiprotons emerging forward in the lab frame due to . The yield is low, typically on the order of 10^{-5} antiprotons per incident proton, necessitating stochastic cooling and accumulation in storage rings like the Antiproton Accumulator to build usable fluxes. Following production, antiprotons with momenta around 3.5 GeV/c are captured by magnetic horns and decelerated in facilities such as CERN's Antiproton Decelerator (AD), which reduces their energy to about 5.3 MeV, or further to 100 keV in the ELENA extension for low-energy experiments. This deceleration enhances trapping efficiency for studies, as slower antiprotons are easier to confine in Penning traps or mix with positrons to form . Historically, the first antiprotons were produced in 1955 at the Berkeley using a similar proton-target method, confirming Dirac's prediction of . Antiproton annihilation occurs when an collides with a or , converting their combined rest mass of 1876 MeV entirely into carried by other particles, primarily mesons, via processes. In proton- annihilation at rest, the dominant channels produce an average of approximately 3 charged pions (\pi^\pm) and 2 neutral pions (\pi^0), with rarer contributions from kaons (about 6% branching ratio) and \eta mesons (about 7%), releasing pions with an average of 230 MeV. The process typically initiates at the nuclear surface, about 1 fm inside the , where the antiproton is captured in an atomic orbit before annihilating on a , often leading to final-state interactions that can eject nucleons or cause breakup. For neutron annihilation, the pion multiplicity shifts to roughly 1 \pi^+ and 2 \pi^- per event, reflecting isospin differences, while in heavier nuclei, the annihilation energy can excite the residual nucleus, producing light ions or "prongs" with multiplicities varying by target material—higher in light elements like than in heavy ones like gold. Experimental measurements from low-energy antiproton beams, such as those at CERN's LEAR and AD, confirm these branching ratios through tracking detectors, revealing discrepancies with simulations like that overestimate prong production by up to 4 times in some cases. Annihilation products, including gamma rays from \pi^0 decays, provide signatures for vertex reconstruction in experiments probing or gravitational effects on .

References

  1. [1]
  2. [2]
    proton mass - CODATA Value
    proton mass $m_{\rm p}$ ; Numerical value, 1.672 621 925 95 x 10-27 kg ; Standard uncertainty, 0.000 000 000 52 x 10-27 kg.
  3. [3]
    What is an atom ? | Nuclear Regulatory Commission
    There are three subatomic particles: protons, neutrons and electrons. Two of the subatomic particles have electrical charges: protons have a positive charge ...
  4. [4]
    1.9: The Properties of Protons, Neutrons, and Electrons
    Electrons are a type of subatomic particle with a negative charge. · Protons are a type of subatomic particle with a positive charge. · Neutrons are a type of ...
  5. [5]
    Protons and neutrons - HyperPhysics Concepts
    Along with neutrons, protons make up the nucleus, held together by the strong force. The proton is a baryon and is considered to be composed of two up quarks ...Missing: subatomic definition
  6. [6]
    For Protons and Neutrons, Things Aren't the Same Inside Nuclei
    Nov 17, 2022 · Nuclear physicists call this difference “the EMC effect.” Each proton is made of three quarks, with two called up quarks and one called a down ...
  7. [7]
    Ernest Rutherford
    Shortly after the World War I, in 1920, Rutherford proposed the name proton for the positively charged particles in the nucleus of an atom. Proposing the ...
  8. [8]
    Manhattan Project: Exploring the Atom, 1919-1932 - OSTI.GOV
    Rutherford found that nitrogen nuclei ejected what he suspected was a hydrogen atom when bombarded with energetic α (alpha) particles.
  9. [9]
    SpinQuest: Putting together the proton spin puzzle - Newsroom
    Mar 10, 2021 · Protons have a spin of ½ and two orientations: spin up and spin down. Protons are composed of even smaller particles called quarks and gluons, ...
  10. [10]
    [PDF] 15. Quark Model - Particle Data Group
    May 31, 2024 · The quark model describes the properties of low-lying bound states of QCD, using the language of quarks and antiquarks, and is older than QCD.
  11. [11]
    [PDF] 18. Structure Functions - Particle Data Group
    Dec 1, 2023 · The double-differential cross section for deep inelastic scattering can be expressed in terms of kinematic variables in several ways. d2σ dx dy.
  12. [12]
    [PDF] 9. Quantum Chromodynamics - Particle Data Group
    May 31, 2024 · The ψq,a are quark-field spinors for a quark of flavor q and mass mq, with a color-index a that runs from a = 1 to Nc = 3, i.e. quarks come in ...Missing: binding | Show results with:binding
  13. [13]
    [PDF] Understanding QCD: From Quarks and Gluons to Hadrons
    → Quarks carry only half of the proton's momentum! The rest? Is carried by the so-called gluons – particles responsible for the interaction between quarks.
  14. [14]
    [PDF] Physics of the Nucleon Sea Quark Distributions - arXiv
    In this picture, the proton has three valence quarks, two up, or u, quarks ... in the constituent quark model, the proton is a uud state where the valence quarks ...Missing: composition | Show results with:composition
  15. [15]
    [PDF] The Proton Spin and Flavor Structure in the Chiral Quark Model - arXiv
    Quark contributions to the proton spin Because it is antisymmetric under the interchange of quark color indices, the baryon wavefunction must be symmetric ...
  16. [16]
    [PDF] 1. Physical Constants - Particle Data Group
    electron mass me. 0.510 998 950 00(15) MeV/c2 = 9.109 383 7015(28)×10−31 kg. 0.30 proton mass mp. 938.272 088 16(29) MeV/c2 = 1.672 621 923 69(51)×10−27 kg.Missing: rest | Show results with:rest
  17. [17]
    [2211.09905] Origin of the Proton Mass - arXiv
    Nov 17, 2022 · This contribution sketches forty years of developments in QCD, which suggest a solution to the puzzle, and highlight some of the experiments that can validate ...
  18. [18]
    proton rms charge radius - CODATA Value
    proton rms charge radius $r_{\rm p}$. Numerical value, 8.4075 x 10-16 m. Standard uncertainty, 0.0064 x 10-16 m. Relative standard uncertainty, 7.6 x 10-4.
  19. [19]
    A measurement of the atomic hydrogen Lamb shift and the proton ...
    Sep 6, 2019 · Our measurement determines the proton radius to be rp = 0.833 femtometers, with an uncertainty of ±0.010 femtometers. This electron-based ...
  20. [20]
    Phys. Rev. D 102, 112011 (2020) - Search for proton decay via $p ...
    Dec 22, 2020 · Search for proton decay via p→e+π0 and p→μ+π0 with an enlarged fiducial volume in Super-Kamiokande I-IV. A. Takenaka et al.
  21. [21]
    Phys. Rev. D 106, 072003 (2022) - Search for proton decay via $p ...
    Oct 10, 2022 · Search for proton decay via p→μ+K0 in 0.37 megaton-years exposure of Super-Kamiokande. R. Matsumoto et al. (Super-Kamiokande Collaboration).
  22. [22]
    The Discovery of the Electron (JJ Thomson) - Purdue University
    At the time Thomson proposed this model, evidence for the existence of positively charged particles was available from cathode-ray tube experiments. fig6_5b ...
  23. [23]
    May, 1911: Rutherford and the Discovery of the Atomic Nucleus
    May 1, 2006 · Rutherford's explanation, which he published in May 1911, was that the scattering was caused by a hard, dense core at the center of the atom–the nucleus.
  24. [24]
    Rutherford, transmutation and the proton - CERN Courier
    May 8, 2019 · The events leading to Ernest Rutherford's discovery of the proton, published in 1919. Nuclear giant Ernest Rutherford in Canada in 1907.
  25. [25]
    [PDF] Discovery of proton - UF Physics Department
    Discovery of proton. 1896. Rutherford arrives to the Cavendish Lab from New Zeeland. Works on conductivity of gases exposed to X-rays and radioactivity ...
  26. [26]
    [PDF] Gell-Mann.pdf
    PHYSICS LETTERS. 1 February 1964. A SCHEMATIC. MODEL OF BARYONS AND MESONS. M. GELL- MANN. California Institute of Technology, Pasadena, California. Received 4 ...
  27. [27]
  28. [28]
    [PDF] DEEP INELASTIC SCATTERING - Nobel Prize
    At the time the MIT-SLAC inelastic experiments started in 1968, there was ... J. I. Friedman, Deep Inelastic Scattering: Comparisons with the Quark Model.
  29. [29]
    Deep inelastic scattering: Experiments on the proton and the ...
    Jul 1, 1991 · Deep inelastic scattering: Experiments on the proton and the observation of scaling. Henry W. Kendall.
  30. [30]
    Ultraviolet Behavior of Non-Abelian Gauge Theories
    Jun 25, 1975 · It is shown that a wide class of non-Abelian gauge theories have, up to calculable loga- rithmic corrections, free-field-theory asymptotic ...
  31. [31]
    Reliable Perturbative Results for Strong Interactions?
    Jun 25, 1973 · Rev. Lett. 28,. 1494 (1972); S. Weinberg, Phys, Rev. D 5, 1962 (1972).
  32. [32]
    [PDF] 17. Lattice Quantum Chromodynamics - Particle Data Group
    Jun 1, 2020 · It is the purpose of this review to provide an outline of the methods of LQCD, with particular focus on applications to particle physics, and an ...
  33. [33]
    DOE Explains...Protons - Department of Energy
    The proton is a subatomic particle with a positive electrical charge. They are found in every atomic nucleus of every element.
  34. [34]
    The Nucleus
    The term nucleon is used for either a proton or a neutron. The simplest nucleus is that of hydrogen, which is just a single proton, while the largest nucleus ...
  35. [35]
    Why doesn't the nucleus fly apart? | Science Guys - Union University
    When the strong force gets its claws on a proton, it pulls the proton firmly to the nucleus. The strong force is approximately 100 times larger than the Coulomb ...
  36. [36]
  37. [37]
    Isotope Basics | NIDC
    Isotopes are atoms of the same element that have the same number of protons (i.e., atomic number, "Z") but a different number of neutrons, meaning that their ...
  38. [38]
    Theoretical Models
    Aug 9, 2000 · The Liquid Drop Model treats the nucleus as a liquid. Nuclear properties, such as the binding energy, are described in terms of volume energy, ...
  39. [39]
    Liquid Drop Model of Nucleus - HyperPhysics
    A liquid drop model of the nucleus would take into account the fact that the forces on the nucleons on the surface are different from those on nucleons on the ...
  40. [40]
    Big Bang Nucleosynthesis
    The Big Bang Nucleosynthesis theory predicts that roughly 25% the mass of the Universe consists of Helium. It also predicts about 0.01% deuterium, and even ...
  41. [41]
    phy213 - the physics of stellar interiors - hydrogen and helium burning
    The PP chain divides into three main branches, which are called the PPI, PPII and PPIII chains. ... PPII chain or captures another proton to form 8B in the PPIII ...
  42. [42]
    [PDF] Stellar Nuclear Fusion Nuclear Reactions - Loren Anderson
    Below I'll describe common pathways for the fusion reactions. The proton-proton chain is the most common fusion sequence in most stars. ... Figure 3: The PPI ...
  43. [43]
    What are Cosmic Rays? | ICRR
    Feb 12, 2019 · About 90 percent of the cosmic rays are hydrogen nuclei―namely protons, 9 percent are helium nuclei, and the rest are heavier nuclei and ...
  44. [44]
    Cosmic rays: particles from outer space - CERN
    The energies of the primary cosmic rays range from around 1 GeV – the energy of a relatively small particle accelerator – to as much as 108 TeV, far higher ...
  45. [45]
    [PDF] 34. Passage of Particles Through Matter
    Jun 1, 2020 · Electron (positron) scattering is considered as ionization when the energy loss per collision is below 0.255 MeV, and as Møller (Bhabha) ...Missing: ordinary | Show results with:ordinary
  46. [46]
    [PDF] Radiotherapy Proton Interactions in Matter - arXiv
    Mar 30, 2018 · Abstract. Radiotherapy protons interact with matter in three ways. Multiple collisions with atomic electrons cause them to lose energy and ...
  47. [47]
    [PDF] Particle Interactions with Matter
    Elastic scattering processes (nuclear elastic or Coulomb scattering) and energy loss processes (ionization) do not affect proton survival and are therefore ...
  48. [48]
    [PDF] Rutherford Scattering - MIT
    We shall now derive the differential cross section do/dN(0, 4), for an alpha particle scattered by the Coulomb field of the nucleus into the ele- ment of solid ...
  49. [49]
    Stopping-Power & Range Tables for Electrons, Protons, and Helium ...
    Stopping-power and range tables can be calculated for electrons in any user-specified material and for protons and helium ions in 74 materials. Update. The data ...
  50. [50]
    [PDF] Classification of the LHC BLM Ionization Chamber
    The main detector type is an ionization chamber with ... Table 2: Range of correction factors due to space charge for all detectors depending on proton beam ...
  51. [51]
    [PDF] Characterization Method of Ionization Chamber Type Detectors
    Alternatively, ionization chambers can be sent to National Metrology Institutes around the world similar to NIST. Transferring calibrations between instruments.
  52. [52]
    [PDF] p +) Status: ∗∗∗∗ - Particle Data Group
    Most measurements of the radius of the proton involve electron-proton interactions, the most recent of which is the electron scattering measurement rp ...
  53. [53]
    Nuclear Forces - Scholarpedia
    Sep 14, 2014 · Yukawa constructed his theory in analogy to the theory of the electromagnetic interaction where the exchange of a (massless) photon is the cause ...Missing: screening | Show results with:screening
  54. [54]
    Fundamental Forces - HyperPhysics Concepts
    One of the four fundamental forces, the electromagnetic force manifests itself through the forces between charges (Coulomb's Law) and the magnetic force.
  55. [55]
    2.4: Nuclei of Atoms - Chemistry LibreTexts
    Apr 9, 2022 · For example, hydrogen has the atomic number 1; all hydrogen atoms have 1 proton in their nuclei. Helium has the atomic number 2; all helium ...
  56. [56]
    The Hydronium Ion - Chemistry LibreTexts
    Jan 29, 2023 · The hydronium ion has a trigonal pyramidal geometry and is composed of three hydrogen atoms and one oxygen atom.Missing: radius | Show results with:radius
  57. [57]
    The pH Scale - Chemistry LibreTexts
    Jan 29, 2023 · For example, at a pH of zero the hydronium ion concentration is one molar, while at pH 14 the hydroxide ion concentration is one molar.Missing: credible | Show results with:credible<|control11|><|separator|>
  58. [58]
    Brønsted Concept of Acids and Bases - Chemistry LibreTexts
    Jan 29, 2023 · Therefore, HCl is a Brønsted-Lowry acid (donates a proton) while the ammonia is a Brønsted-Lowry base (accepts a proton).
  59. [59]
    Brønsted-Lowry acids and bases (article) - Khan Academy
    A Brønsted-Lowry acid is any species that can donate a proton, H ... Because water accepts a proton from HCl ‍ , water is acting as a Brønsted-Lowry base.
  60. [60]
    Tunneling in proton-transfer reactions in solution | Chemical Reviews
    This article is cited by 144 publications. Surajit Kalita, David Danovich, Sason Shaik. Local Electric Fields Originate Unusual Kinetic Isotope Effects in ...Missing: credible | Show results with:credible
  61. [61]
    Large kinetic isotope effects in enzymatic proton transfer and the ...
    The subject of this paper is the study of the anomalous kinetic isotope effect (KIE) indicative of tunneling that recently has been observed in certain ...
  62. [62]
    Hydrogen Tunneling Links Protein Dynamics to Enzyme Catalysis
    We show how the quantum mechanical hydrogen tunneling associated with enzymatic C–H bond cleavage provides a unique window into the necessity of protein ...
  63. [63]
    [PDF] Basics of NMR Spectroscopy - UConn Health
    Nov 29, 2016 · A nuclei with spin can be thought of as if it were a tiny bar magnet. • The gyromagnetic ratio (γ) is an empirically determined constant.
  64. [64]
    [PDF] Basic NMR Concepts:
    This guide is intended to highlight the most pertinent aspects of practical NMR spectroscopy. It is not meant to replace any formal training of the theory of ...Missing: fundamentals | Show results with:fundamentals
  65. [65]
    [PDF] Introduction to Nuclear Magnetic Resonance Spectroscopy
    600 MHz is the frequency at which the proton (1H) nucleus spin resonates – in a magnet of this strength. The magnet is superconducting, always charged, but not ...Missing: fundamentals | Show results with:fundamentals
  66. [66]
    Proton therapy in clinical practice - PMC - PubMed Central
    The increment of proton dose at a specified area is referred to as the Bragg Peak[9]. This ability allows proton therapy to spare healthy tissue and have a ...
  67. [67]
    History and Heritage of Proton-Beam Therapy: Robert R. Wilson
    His article got the attention of radiation therapists and other physicians. Proton beams began to be used therapeutically in the mid-1950s; the Proton Treatment ...Missing: early | Show results with:early
  68. [68]
    Proton Therapy – Present and Future - PMC - PubMed Central
    Protons, accelerated to therapeutic energies ranging from 70 to 250 MeV, typically with a cyclotron or a synchrotron, are transported to the treatment room ...
  69. [69]
    More than Five Decades of Proton Therapy: A Bibliometric Overview ...
    Nov 23, 2023 · Nowadays, over 100 proton centers are in operation worldwide, and more than 60 are under construction or planned for a total of more than 311, ...3. Results · 3.4. Themes And Trends · 4. Discussion
  70. [70]
    Proton induced DNA double strand breaks at the Bragg peak
    Aug 4, 2022 · They are plotted as a function of LET, which shows value of 1.1 keV/µm at 2 cm and 22.6 keV/µm at 9.4 cm depth. This quantity is calculated as ...
  71. [71]
    Physical and Biological Characteristics of Particle Therapy for ... - NIH
    The characteristic feature of particle therapy is the “Bragg peak,” a steep and localized peak of dose, that enables precise delivery of the radiation dose ...
  72. [72]
    [PDF] Risk of Acute Radiation Syndromes due to Solar Particle Events
    Crew health and performance may be impacted by a major solar particle event (SPE), multiple. SPEs, or the cumulative effect of galactic cosmic rays (GCR) ...
  73. [73]
    Acute effects of solar particle event radiation - PMC - NIH
    A major solar particle event (SPE) may place astronauts at significant risk for the acute radiation syndrome (ARS), which may be exacerbated when combined ...
  74. [74]
    The magnitude and effects of extreme solar particle events
    The SPE of August 1972 is often cited as a prototypical worst- case SPE which could result in severe health risks to astronauts on EVA or even death. It ...
  75. [75]
    Report of the AAPM TG‐256 on the relative biological effectiveness ...
    Jan 19, 2019 · The biological effectiveness of proton beams relative to photon beams in radiation therapy has been taken to be 1.1 throughout the history ...
  76. [76]
    New Recommendations for Occupational Radiation Protection
    The International Commission on Radiological Protection recommends that effective dose be limited to 20 mSv/year averaged over 5 consecutive years, with no ...
  77. [77]
    [PDF] Galactic and Solar Cosmic Ray Shielding in Deep Space
    In this context, polyethylene is examined as a potentially useful material and demonstrates important advantages as an alternative to aluminum construction.
  78. [78]
    [PDF] Early Results from the Advanced Radiation Protection Thick GCR ...
    Polyethylene has a large hydrogen content and therefore does a better job of attenuating neutrons and protons, thereby minimizing the buildup of light ...
  79. [79]
    Observation of Antiprotons | Phys. Rev.
    Observation of Antiprotons. Owen Chamberlain, Emilio Segrè, Clyde Wiegand, and Thomas Ypsilantis. Radiation Laboratory, Department of Physics, University of ...
  80. [80]
    The Nobel Prize in Physics 1959 - NobelPrize.org
    The Nobel Prize in Physics 1959 was awarded jointly to Emilio Gino Segrè and Owen Chamberlain for their discovery of the antiproton.
  81. [81]
    A parts-per-billion measurement of the antiproton magnetic moment
    Oct 19, 2017 · The magnetic moment of the antiproton is measured at the parts-per-billion level, improving on previous measurements by a factor of about ...
  82. [82]
    Antiproton Source | ALPHA Experiment - CERN
    To produce antiprotons, we smash the proton beam into a proton rich metal target (iridium). In this collision, secondary particles get created, including among ...
  83. [83]
    CERN antiproton target: Hydrocode analysis of its core material ...
    Jul 15, 2016 · Antiprotons are produced at CERN by colliding a 2 6 G e V / c proton beam with a fixed target made of a 3 mm diameter, 55 mm length iridium core ...
  84. [84]
  85. [85]
    Antiproton annihilation at rest in thin solid targets and comparison ...
    Nov 17, 2024 · Antiproton ( ) annihilation on a nucleus (A) is a particular nuclear reaction in which part of the released energy is transferred to the nucleus ...
  86. [86]
    Antimatter
    When a particle and its antiparticle meet, they annihilate into pure energy. During an annihilation, a matter and antimatter particle interact with each other, ...