Fact-checked by Grok 2 weeks ago

Atomic nucleus

The atomic nucleus is the small, dense, positively charged core at the center of an , composed primarily of protons and neutrons tightly bound together. It was discovered in 1911 by through his gold foil experiment, which revealed that most alpha particles fired at a thin gold foil passed through undeflected, indicating the atom's mass and positive charge are concentrated in a tiny central region. The nucleus accounts for over 99.9% of the 's mass but occupies only about 10^{-15} of its volume, with typical diameters ranging from 2 to 15 femtometers (10^{-15} meters), compared to the atom's diameter of around 100 picometers (10^{-10} meters). Protons, which carry a positive electric charge equal in magnitude to that of an , determine the of an and thus its chemical identity, while neutrons, which are electrically neutral, contribute to the 's stability without altering the charge. The number of protons (Z) and neutrons (N) together defines the (A = Z + N), and variations in neutron count lead to isotopes of the same , such as (6 protons, 6 neutrons) and (6 protons, 8 neutrons)./03%3A_Atomic_Structure/3.05%3A_The_Atomic_Nucleus) The protons' mutual electrostatic repulsion is overcome by the , a mediated by gluons between quarks within the protons and neutrons (collectively called nucleons), which acts at short ranges of about 1-2 femtometers to bind the . This force is approximately 100 times stronger than the electromagnetic force but drops off rapidly beyond the nuclear scale. Nuclear stability depends on the balance between protons and neutrons; lighter nuclei typically have roughly equal numbers, while heavier ones require more neutrons to dilute proton repulsion, with the most stable configurations around iron-56. Unstable nuclei undergo , emitting alpha particles ( nuclei), beta particles (electrons or positrons), or gamma rays to reach a more stable state, processes central to and applications like energy production and . The liquid drop model and are key theoretical frameworks for understanding nuclear structure, treating the nucleus as a semi-classical droplet or quantum system with energy levels akin to electrons in atoms. Ongoing research explores nuclear shapes, excitations, and reactions, revealing complexities such as proton-neutron pairing and the role of the weak nuclear force in .

Historical Development

Early Concepts and Observations

The concept of the atomic nucleus emerged from a long philosophical and scientific tradition that initially viewed atoms as indivisible units without internal structure. In the 5th century BCE, the Greek philosophers and proposed an positing that all matter consists of eternal, indivisible particles called atomos (meaning "uncuttable"), which differ in shape, size, and arrangement to form the variety of substances observed in nature. This idea served as a philosophical precursor to later atomic models but lacked any notion of subatomic components or a centralized , emphasizing instead atoms as the fundamental building blocks moving in a void. By the early 19th century, experimental chemistry revived and refined atomic ideas, though still without recognizing nuclear structure. John Dalton's , outlined in his 1808 publication A New System of Chemical Philosophy, asserted that elements consist of identical, indivisible atoms that combine in fixed ratios to form compounds, explaining laws such as definite proportions and multiple proportions. Dalton's model portrayed atoms as solid, uniform , providing a quantitative foundation for chemistry but assuming no internal differentiation. Toward the end of the century, J.J. Thomson's discovery of the in 1897 prompted a revision, leading to his 1904 "plum pudding" model, which depicted the atom as a of uniformly distributed positive charge embedding negatively charged electrons like plums in pudding to maintain overall neutrality. This model accounted for electrical stability but implied a diffuse, non-concentrated positive charge throughout the atom. The late 19th century also brought indirect observations hinting at atomic instability through , challenging the indivisibility of atoms. In 1896, discovered that salts emit invisible rays capable of penetrating opaque materials and exposing photographic plates, a phenomenon he termed "uranium rays," independent of external stimulation like light. This finding revealed spontaneous atomic emissions, suggesting internal energy sources within atoms. Building on Becquerel's work, Marie and Pierre Curie isolated two highly radioactive elements from pitchblende ore: in July 1898, which was about 400 times more active than , and in December 1898, exhibiting even greater intensity. Their extractions demonstrated that radioactivity arises from specific atomic species, implying concentrated sources of emission within matter. These developments influenced , who, as director of the Manchester laboratory, sought to probe atomic structure using alpha particles from radioactive sources. In 1908, Rutherford suggested to his assistant that they investigate whether alpha particles could scatter at large angles through thin metal foils, expecting only minor deflections consistent with Thomson's diffuse charge model. This setup, later executed with in 1909, marked a pivotal shift toward direct experimental interrogation of the atom's interior.

Key Experiments and Discoveries

In 1909, and conducted a series of experiments at the , directing a beam of alpha particles from a radioactive source onto thin foil and observing their patterns using a fluorescent screen and . They found that while most alpha particles passed through the foil undeflected, a small fraction underwent large-angle deflections, with some rebounding nearly 180 degrees back toward the source, and that the number of particles deflected by angles greater than 90 degrees was approximately 1 in 8,000 incident particles. Further refinements in 1913 involved measuring the distribution of angles more precisely and deriving the laws governing large-angle deflections, which supported Rutherford's theoretical predictions. These unexpected results, which contradicted the prevailing plum pudding model of the atom, prompted Ernest Rutherford to analyze the data theoretically in 1911. Rutherford proposed that the observed large-angle scatterings could only be explained by the repulsion between the positively charged alpha particles and a small, dense, positively charged core within the atom, which he termed the nucleus; this core was estimated to occupy less than 1/10,000th of the atom's volume yet contain nearly all its mass. The model predicted scattering probabilities that matched the experimental observations quantitatively, establishing the nuclear structure as the foundation of atomic architecture. Building on studies of chains, introduced the concept of isotopes in 1913 to account for chemically identical elements exhibiting different radioactive properties and atomic masses. By examining sequences where decayed through multiple steps to stable lead, Soddy observed that intermediate products, such as certain emanations and ionium, shared identical chemical behaviors despite varying atomic weights, implying non-integral mass contributions from nuclear constituents. This insight resolved discrepancies in schemes and highlighted the nucleus's role in determining both chemical identity and isotopic variation. In 1932, resolved longstanding puzzles about atomic mass by discovering the through irradiation experiments at the . He bombarded with alpha particles from , producing a highly penetrating neutral radiation that, unlike gamma rays, ejected protons from with energies up to 5.7 MeV, indicating collisions with particles of mass roughly equal to the proton. Interpreting this as evidence for an uncharged nuclear particle—the —Chadwick's work explained the integer mass excesses in light elements and completed the basic composition of the as protons and neutrons.

Etymology and Terminology

The term "" originates from the Latin nux, meaning "," evoking the idea of a hard central . introduced it in 1911 to describe the dense, positively charged core of the atom, inferred from the scattering of alpha particles in his gold foil experiment. Rutherford further advanced nuclear terminology in 1920 with the term "," derived from the Greek prōtos ("first"), to designate the positively charged particle constituting the nucleus and serving as a basic building block of all atomic nuclei. This naming reflected its fundamental role, building on earlier observations of hydrogen ions ejected from atomic collisions. In 1932, discovered a neutral particle of approximately the proton's mass within the nucleus and named it the "," a term emphasizing its lack of and paralleling the suffixes of "proton" and "." The nomenclature highlighted the particle's role in explaining nuclear mass without additional charge. As progressed, terminology evolved to distinguish specific nuclear configurations; for instance, the term "" was coined by Truman P. Kohman in 1947 to refer to a particular species of defined by its proton and neutron numbers, facilitating precise descriptions of isotopes and their properties.

Basic Principles

Definition and Role in Atoms

The atomic nucleus is the dense central core of an atom, composed of protons and neutrons collectively known as nucleons. It was discovered by in 1911 through gold foil scattering experiments. The nucleus contains nearly all of an atom's mass, accounting for more than 99.9994% of the total despite occupying an extremely small volume. The number of protons in the defines the Z, which uniquely identifies the of the atom. For instance, all atoms have Z = 1, while carbon atoms have Z = 6. This proton count determines the element's position in the periodic table and its fundamental chemical identity. The positive charge of the nucleus, equal to Z times the , exerts a attraction on the surrounding , influencing their in atomic orbitals. This nuclear charge shapes the structure, which in turn governs the atom's chemical properties and bonding behavior. Elements with similar exhibit , such as and reactivity, directly tied to the underlying nuclear charge. The nucleus serves as the source of atomic stability by providing the centralized positive charge that binds the negatively charged electrons, countering their mutual electrostatic repulsion and forming a cohesive atomic structure. Without this binding, the electrons would not maintain their orbits, rendering the atom unstable.

Composition and Structure

The atomic nucleus is composed of two types of fundamental particles known as nucleons: protons and neutrons. Protons carry a positive equal to the e (approximately $1.602 \times 10^{-19} C) and have a rest mass of about $1.67 \times 10^{-27} . Neutrons, in contrast, are electrically neutral and possess a nearly identical rest mass of approximately $1.67 \times 10^{-27} . The total number of nucleons in a nucleus defines its mass number A, which is the sum of the proton number Z (also called the atomic number) and the neutron number N, expressed as A = Z + N. This integer A approximates the atomic mass in unified atomic mass units (u), though the actual mass is slightly less due to binding effects. Nuclei with the same Z but different N (and thus different A) are classified as isotopes of the same chemical element, such as the stable isotopes ^{12}C (Z=6, N=6) and ^{13}C (Z=6, N=7). Nuclei sharing the same A but differing in Z and N are isobars, like ^{14}C (Z=6, N=8) and ^{14}N (Z=7, N=7), which often exhibit beta decay pathways connecting them. Isotones, meanwhile, have the same N but different Z and A, such as ^{13}C (Z=6, N=7) and ^{14}N (Z=7, N=7). In terms of internal arrangement, protons and neutrons are bound together within a compact volume, forming a dense core at the atom's center. At the scale of , these nucleons are treated as the primary building blocks without further subdivision; deeper substructure involving quarks and gluons falls under the domain of . This composite structure underpins the nucleus's role in determining an atom's identity and stability.

Size, Shape, and Density

The R of an atomic nucleus is empirically described by the formula R \approx 1.2 \times 10^{-15} A^{1/3} meters (or 1.2 A^{1/3}, where 1 = $10^{-15} m and A is the ), derived from high-energy scattering experiments that probe the nuclear charge distribution. This relation arises because nuclear volume scales proportionally with A, assuming a roughly constant of nucleons. This empirical relation provides reasonable estimates for nuclei with larger A; for example, for (A = 238), R \approx 7.4 fm. For the proton (hydrogen-1 nucleus, A = 1), the formula gives 1.2 fm, but the measured root-mean-square is 0.841 fm. In stark contrast, the radius of an entire is typically on the order of $10^{-10} m (or 100,000 fm), highlighting the nucleus's extreme compactness within the atomic structure./University_Physics_III_-Optics_and_Modern_Physics(OpenStax)/11%3A_Particle_Physics_and_Cosmology/11.06%3A_The_Nucleus) Nuclear shapes deviate from perfect sphericity in many cases, with even-even nuclei (equal numbers of protons and neutrons) often approximating spheres in their ground states due to effects that favor symmetric configurations. However, numerous nuclei display deformations, appearing prolate (elongated like a ) or (flattened like a ), as revealed by the angular dependence of elastic cross-sections, which show non-spherical form factors beyond the first minimum. These deformations are more pronounced in transitional regions of the nuclear chart and in nuclei with odd nucleon numbers, where single-particle asymmetries disrupt spherical symmetry. The of is remarkably uniform and independent of A, saturating at approximately $2.3 \times 10^{17} kg/m³ (or $0.17 s per fm³), a value that reflects the balance of the strong nuclear force at short ranges. This constant " density" implies that all nuclei, regardless of size, have a similar central packing of protons and neutrons, with the surface region accounting for any minor variations. The mass is calculated from the average mass (\approx 1.67 \times 10^{-27} kg) divided by the nuclear volume derived from the formula, confirming the near-invariance across isotopes. Measurements of , , and rely on techniques that probe the charge or , such as high-energy , which extends Rutherford's method by resolving finite-size effects through patterns in the differential cross-section. Complementary data come from muonic atom , where orbital muons (207 times heavier than electrons) produce transitions highly sensitive to the nuclear radius via the , and from muon capture rates, which correlate with the overlap of the muon wavefunction and nuclear volume. These methods provide precise root-mean-square charge radii, typically agreeing within 1-2% for nuclei.

Nuclear Forces and Interactions

Strong Nuclear Force

The , also known as the , is the fundamental force responsible for binding protons and neutrons together within the atomic nucleus, overcoming the electromagnetic repulsion between protons. This force arises as a effect of the described by (QCD), where quarks within nucleons exchange gluons that carry , leading to an effective attraction between color-neutral nucleons. In QCD, the strong force between quarks is mediated by massless gluons, which themselves participate in interactions due to , resulting in the short-range between composite nucleons. Key properties of the strong nuclear force include its extremely short range of approximately 1 to 2 , confining its influence to distances comparable to the size of the . It is attractive between and acts independently of , meaning the force is the same for proton-proton (pp), neutron-neutron (nn), and proton-neutron (pn) pairs—a property known as . This charge independence stems from the underlying of the strong interaction, which treats protons and neutrons as two states of the same isospin doublet (), with the up and down quarks exhibiting approximate . The force's strength is about 100 times greater than the electromagnetic force at nuclear scales, enabling it to dominate and stabilize multi-proton . At the phenomenological level, the strong nuclear force between nucleons is primarily mediated by the exchange of pions, the lightest mesons, which act as virtual particles in the interaction. This one-pion exchange (OPE) mechanism gives rise to the , a form originally proposed by in 1935 to describe the . The potential is expressed as: V(r) \approx -\frac{g^2}{4\pi} \frac{e^{-\mu r}}{r}, where g is the pion-nucleon coupling constant, r is the distance between nucleons, and \mu \approx m_\pi c / \hbar is the inverse range parameter set by the pion rest mass m_\pi \approx 140 MeV/c², yielding a range of about 1.4 fm. This exponential decay ensures the force's confinement to nuclear dimensions, dropping rapidly beyond 2-3 fm.

Electromagnetic and Weak Forces in Nuclei

The electromagnetic force in the atomic nucleus arises from the electrostatic repulsion between positively charged protons, which opposes the attractive strong nuclear force and influences nuclear stability. This repulsion creates the , a hurdle that must be overcome for protons to approach closely enough for nuclear interactions. The height of the Coulomb barrier for two interacting nuclei is described by the formula V_c = \frac{Z_1 Z_2 e^2}{4 \pi \epsilon_0 r}, where Z_1 and Z_2 are the proton numbers of the nuclei, e is the elementary charge, \epsilon_0 is the permittivity of free space, and r is the center-to-center separation distance. In light nuclei, this barrier is particularly limiting, as the long-range nature of the electromagnetic interaction requires the short-range strong force to dominate at very small distances to achieve binding, restricting the number of protons relative to neutrons for overall stability. The strong nuclear force serves as the primary counterbalance to this repulsion, enabling cohesive nuclear structures despite the perturbative electromagnetic effects. The Coulomb interaction also directly impacts nuclear binding energy, as captured in the semi-empirical mass formula through a negative Coulomb term, approximately -a_c Z(Z-1)/A^{1/3}, where a_c \approx 0.7 MeV is an empirical coefficient, Z is the proton number, and A is the mass number. This term accounts for the pairwise repulsion among protons, which decreases the total binding energy and favors neutron-rich compositions in heavier nuclei to minimize electrostatic penalties./01%3A_Introduction_to_Nuclear_Physics/1.02%3A_Binding_energy_and_Semi-empirical_mass_formula) Without this correction, the formula would overestimate binding for proton-rich systems, highlighting how electromagnetic effects set an upper limit on Z for stable isotopes across the periodic table. In contrast, the weak nuclear force, far feebler than the electromagnetic interaction, governs processes that change particle flavor, such as , exemplified by the transformation n \to p + e^- + \bar{\nu}_e. Its relative strength is characterized by a roughly $10^{-6} times that of the strong force, compared to the electromagnetic force's \alpha \approx 1/137, making it the weakest of the fundamental interactions yet essential for enabling nuclear transmutations like those in . Evidence for the weak force's unique properties, including parity violation, came from the Wu experiment, which demonstrated asymmetric in the of polarized ^{60}Co nuclei, confirming that weak interactions do not conserve mirror symmetry. Weak interactions are particularly influential in odd-A nuclei, where an unpaired disrupts pairing correlations, leading to elevated rates and reduced stability compared to even-A counterparts. In these systems, the weak force facilitates transitions between neutron- and proton-rich states, contributing to the observed prevalence of beta-unstable odd-mass isotopes in . This role underscores the weak force's importance in shaping isotopic distributions, as it allows odd-A nuclei to evolve toward more stable configurations through , despite the dominance of stronger forces in binding.

Range Limits and Halo Nuclei

The strong , responsible for protons and s within the atomic , operates over a finite range of approximately 1–2 , beyond which its attractive influence diminishes rapidly. This short range imposes fundamental limits on nuclear , preventing the binding of additional nucleons once a critical imbalance occurs between protons and neutrons. As a result, nuclei with excessive neutrons or protons approach the or proton drip lines—demarcation boundaries where the separation energy of the least-bound valence nucleon approaches zero, rendering the unbound against particle emission. Near these drip lines, certain exotic nuclei develop structures, characterized by one or more loosely bound nucleons whose probability density extends significantly beyond the compact core of tightly bound nucleons, forming a diffuse "." The low of these particles, often on the order of a few hundred keV, allows their wave functions to spread over distances much larger than the core radius (typically 2–3 ), despite the confining short range of the strong force. This configuration arises from the interplay of the strong force's finite reach and the repulsion (for protons) or Pauli exclusion effects, enabling weakly bound states that probe the limits of distribution. A seminal review highlights that such are threshold phenomena, where the proximity to the threshold amplifies the spatial extension of the nucleons. The archetypal two-neutron halo nucleus ^{11}\mathrm{Li}, consisting of a ^{9}\mathrm{Li} core plus two valence neutrons, exemplifies this phenomenon. Discovered in 1985 through measurements of anomalously large interaction cross sections in high-energy collisions at the Bevalac accelerator, ^{11}\mathrm{Li} revealed a matter radius of about 3.5 fm, about 1.5 times larger than that of stable lithium isotopes—attributable to the extremely low two-neutron separation energy of 0.369 MeV. This loose binding results in the valence neutrons orbiting at average distances of 5–6 fm from the core, forming a Borromean structure where no two-body subsystem (core-neutron or neutron-neutron) is bound alone. Other notable examples include ^{6}\mathrm{He}, a two-neutron around an \alpha-particle with of 0.98 MeV, and ^{8}\mathrm{B}, a one-proton with a proton separation of just 0.14 MeV. These structures are probed experimentally via fragmentation and reactions at facilities such as RIKEN's Radioactive , where relativistic beams of halo nuclei are collided with light or heavy targets to measure momentum distributions and decay products. For instance, fragmentation studies of ^{6}\mathrm{He} at 240 MeV/u have confirmed its extended halo through kinematically complete dissociation events, while near-barrier experiments on ^{8}\mathrm{B} with tin targets have elucidated the proton's peripheral dynamics via fragment correlations. The presence of halos leads to enhanced total reaction cross sections, often 2–3 times larger than for non-halo isotopes of similar , due to the increased geometric and probability of peripheral interactions. These challenge standard nuclear models like the independent-particle , which underestimate the extended densities near drip lines, and instead require specialized approaches such as few-body models or effective theories to capture the quantum correlations and low-energy of the valence nucleons.

Theoretical Models

Liquid Drop Model

The liquid drop model conceptualizes the atomic nucleus as a charged droplet of incompressible liquid, where nucleons are bound by short-range forces analogous to molecular interactions in a , with playing a key role in maintaining spherical shape. This macroscopic approach, drawing on the saturation properties of nuclear forces, was formalized by and John A. Wheeler in 1939 to elucidate the dynamics of , treating the nucleus as deformable under while preserving overall volume. Central to the model is the (SEMF), originally proposed by in 1935, which empirically parameterizes the nuclear mass M(A, Z) as a function of the A and Z: M(A, Z) = Z m_p + (A - Z) m_n - a_v A + a_s A^{2/3} + a_c \frac{Z(Z-1)}{A^{1/3}} + a_a \frac{(A - 2Z)^2}{A} \pm \delta Here, m_p and m_n are the free proton and neutron masses, respectively; a_v \approx 15.5 MeV is the volume coefficient; a_s \approx 16.8 MeV is the surface coefficient; a_c \approx 0.72 MeV is the coefficient; a_a \approx 23.3 MeV is the asymmetry coefficient; and \delta is the pairing term (approximately 11–12 MeV/A^{1/2}, zero for odd A, positive for even-even nuclei, and negative for odd-odd nuclei). The is then derived as B(A, Z) = [Z m_p c^2 + (A - Z) m_n c^2 - M(A, Z) c^2], with the negative signs in the mass formula reflecting contributions. The volume term -a_v A captures the bulk binding from the strong nuclear force, assuming constant energy per nucleon for large A due to its short range and saturation. The surface term +a_s A^{2/3} corrects for reduced coordination at the nuclear surface (proportional to surface area $4\pi R^2 \propto A^{2/3}, with radius R \propto A^{1/3}), leading to lower stability for smaller nuclei. The Coulomb term +a_c Z(Z-1)/A^{1/3} accounts for electrostatic repulsion among protons, scaling as Z^2 / R and destabilizing heavier elements. The asymmetry term +a_a (A - 2Z)^2 / A arises from the preference for N \approx Z (where N = A - Z) to minimize kinetic energy differences between proton and neutron Fermi seas under the Pauli principle, despite the charge-independent strong force. The pairing term \pm \delta empirically adjusts for enhanced binding in nuclei with paired nucleons, reflecting cooperative quantum effects. This framework applies the liquid drop analogy to predict energies across the periodic table and compute barriers, estimating the for deforming a into asymmetric fragments until scission, as the distorted shape increases surface and energies while conserving volume. For instance, the model quantifies how absorption in heavy nuclei like lowers the barrier, enabling with modest excitation. The liquid drop model excels in describing average binding energy trends and fission energetics for medium-to-heavy nuclei (A > 50), where macroscopic effects dominate, but it underperforms for light nuclei and near magic numbers due to neglect of microscopic shell structure, resulting in systematic overestimation of masses by up to several MeV.

Shell Model

The nuclear shell model is a quantum mechanical framework that describes the atomic nucleus by treating protons and neutrons (collectively nucleons) as independent particles occupying discrete energy levels, analogous to electrons in atomic orbitals. This independent-particle approximation assumes nucleons move in a central mean-field potential generated by the nucleus as a whole, leading to quantized shell structures that explain periodic variations in nuclear properties. The model incorporates the Pauli exclusion principle, with nucleons filling shells according to their quantum numbers, including orbital angular momentum l, total angular momentum j = l \pm 1/2, and isospin to distinguish protons and neutrons. The shell model originated in 1949 through independent developments by and J. Hans D. Jensen, who built on earlier ideas of orbits but crucially introduced a strong spin-orbit coupling term to the single-particle Hamiltonian. This coupling splits energy levels with the same l but different j, aligning the model with experimental observations of nuclear stability. Mayer's work emphasized closed-shell configurations, while Jensen, along with Otto Haxel and , focused on the implications for magic numbers. Their contributions earned them the 1963 . A hallmark of the shell model is its explanation of magic numbers—specific nucleon counts (2, 8, 20, 28, 50, 82, 126) where nuclei exhibit exceptional stability due to completely filled shells, analogous to in chemistry. These numbers arise from the ordering of single-particle levels influenced by spin-orbit coupling, with filled subshells conferring zero net and enhanced . For instance, doubly magic nuclei like ^{16}O (8 protons, 8 neutrons) and ^{208}Pb (82 protons, 126 neutrons) show closed shells, leading to low reactivity in nuclear reactions and systematic trends in binding energies. The model's Hamiltonian for A nucleons is typically expressed as H = \sum_{i=1}^A \frac{\mathbf{p}_i^2}{2m} + \sum_{i<j} V_{ij}, where the first term is the kinetic energy, and the two-body interaction V_{ij} is approximated in the mean-field limit by a single-particle potential U(r) for each nucleon. A common choice for U(r) is the Woods-Saxon form, U(r) = -\frac{V_0}{1 + \exp\left(\frac{r - R}{a}\right)}, with depth V_0 \approx 50 MeV, nuclear radius R \approx 1.25 A^{1/3} fm, and surface diffuseness a \approx 0.65 fm; a spin-orbit term U_{ls} (\mathbf{l} \cdot \mathbf{s}) is added to reproduce level splittings. This potential yields bound states grouped into major shells separated by ~10-20 MeV, facilitating computational tractability. The shell model successfully predicts ground-state properties, particularly for nuclei near closed shells. For odd-mass nuclei, the spin and parity of the ground state match the unfilled single-particle level's j^\pi, such as j = 5/2^+ for ^{17}O from the $1d_{5/2} orbital. Magnetic moments follow Schmidt lines derived from single-particle operators, capturing trends like the deviation of odd-proton nuclei from the proton g-factor due to orbital contributions, though quenching effects require effective charges for precision. Excited states are described as single-particle or single-hole excitations relative to the core, predicting low-lying spectra in semi-magic nuclei, e.g., the first excited state in ^{15}N as a spin-flip transition. To account for residual nucleon-nucleon interactions beyond the mean field, the shell model is extended using the Hartree-Fock (HF) method, which self-consistently determines the single-particle potential from a realistic two-body force, such as the . HF calculations optimize the Slater determinant wave function to minimize energy, generating deformed or spherical mean fields and improving predictions for binding energies and radii across the periodic table. This approach bridges the independent-particle picture with collective effects in midshell regions.

Collective and Cluster Models

The collective model, pioneered by and in the early 1950s, treats the atomic nucleus as a deformable body capable of undergoing cooperative motions among its constituent nucleons, rather than independent particle movements. This approach unifies aspects of the with single-particle effects, focusing on low-energy excitations where the nucleus behaves like a vibrating or rotating fluid. In spherical nuclei, the model predicts vibrational spectra characterized by quadrupole modes, where the nuclear surface oscillates, leading to evenly spaced energy levels for multiphonon states. For deformed nuclei, particularly even-even isotopes in regions away from closed shells, the collective model describes rotational spectra as sequences of levels with moments of inertia consistent with a rigid rotor, but modulated by the nuclear deformation parameter β. The ground-state band in such nuclei exhibits energy spacings following E_J ≈ (ℏ²/2ℐ) J(J+1), where J is the angular momentum and ℐ is the moment of inertia, successfully reproducing observed spectra in rare-earth and actinide regions. Vibrational and rotational degrees of freedom are coupled in transitional nuclei, with the model incorporating β and γ deformations to capture triaxial shapes. The cluster model complements the collective framework by positing that certain nuclei, especially light ones, can be understood as assemblies of tightly bound subclusters like alpha particles (⁴He nuclei), which act as nearly independent units at low energies due to the short-range nature of the strong force. A prominent example is the carbon-12 nucleus, where the Hoyle state—a resonant 0⁺ excitation at 7.65 MeV crucial for stellar nucleosynthesis—is interpreted as a dilute gas-like configuration of three alpha particles arranged in an equilateral triangle or bent chain, rather than a compact shell-model structure. This state, predicted theoretically to facilitate triple-alpha fusion and observed experimentally, decays predominantly into three alphas, supporting the cluster picture. Algebraic formulations of these ideas, such as the interacting boson model (IBM) developed by Akito Arima and Francesco Iachello in the mid-1970s, map collective and cluster phenomena onto a system of interacting s- and d-bosons (representing nucleon pairs), enabling exact solutions via dynamical symmetries like U(5) for vibrations, O(6) for gamma-unstable deformations, and SU(3) for rotations. The IBM effectively describes spectra and transition rates in medium-to-heavy nuclei, including cluster-like states in light systems through extensions like the alpha-cluster IBM. Applications of collective and cluster models extend to interpreting low-energy excitations, such as the spacing and intensities of rotational bands and vibrational multiplets, as well as higher-energy giant resonances like the isoscalar giant quadrupole resonance, which arises from coherent surface vibrations across the nucleus. Experimental evidence includes enhanced cross-sections in transfer reactions, such as (d,p) deuteron stripping, which preferentially populate collective states in deformed targets, revealing their multi-particle character and deformation-dependent spectroscopic factors. In cluster scenarios, reactions like (α,γ) or breakup processes confirm alpha-substructure in light nuclei. Hybrid approaches integrate collective and cluster dynamics with shell-model orbitals, particularly effective for midshell nuclei where maximum deformation occurs and single-particle effects couple strongly to collective modes. In these regions, such as the rare-earth isotopes near N=90, projected shell-model calculations incorporate collective rotations and vibrations onto shell configurations, accurately predicting band structures and electromagnetic transitions without ad hoc parameters. This synergy, rooted in the original unified model, highlights how collective behaviors emerge from microscopic nucleon interactions in transitional regimes.

Nuclear Stability and Phenomena

Binding Energy and Mass Defect

The binding energy of a nucleus is the energy required to separate it into its constituent protons and neutrons, representing the stability imparted by the strong nuclear force overcoming the repulsive electromagnetic forces between protons. This energy arises from the mass defect, the difference between the mass of the isolated nucleons and the mass of the bound nucleus. The total binding energy B is given by B = \left[ Z m_p + N m_n - M \right] c^2, where Z is the atomic number (number of protons), N is the neutron number, m_p and m_n are the masses of the proton and neutron, M is the mass of the nucleus, and c is the speed of light. The mass defect \Delta m = Z m_p + N m_n - M reflects the conversion of mass into binding energy via Einstein's mass-energy equivalence. To assess nuclear stability across isotopes, the binding energy per nucleon B/A (where A = Z + N is the mass number) is plotted against A, forming a characteristic curve that rises sharply for light nuclei, peaks near iron-56, and gradually declines for heavier elements. Iron-56 exhibits one of the highest values at approximately 8.79 MeV per nucleon, indicating maximal stability per particle. This peak explains the energetics of nuclear fusion for light elements (releasing energy by approaching the peak) and fission for heavy elements (releasing energy by moving toward it). Precise measurements of nuclear masses, essential for calculating binding energies, are obtained through techniques like Penning-trap mass spectrometry, which determines ion cyclotron frequencies in strong magnetic fields to achieve relative uncertainties below $10^{-8}. Additionally, binding energies can be inferred from Q-values of nuclear reactions, where the reaction energy release Q equals the difference in binding energies between products and reactants, derived from measured mass differences. Theoretical predictions of binding energies often rely on the semi-empirical mass formula, which approximates the mass defect based on the liquid drop model by including volume, surface, Coulomb, asymmetry, and pairing terms. This formula, originally formulated by Carl Friedrich von Weizsäcker in 1935, provides accurate estimates for stable nuclei with deviations typically under 1%.

Nuclear Reactions and Fission/Fusion

Nuclear reactions involve interactions between atomic nuclei that alter their composition or internal energy states, often releasing or absorbing energy based on differences in nuclear binding. These processes are classified into several types, including elastic scattering, where the incident particle bounces off the target nucleus without changing its internal state, conserving both kinetic energy and momentum in the center-of-mass frame. Inelastic scattering, by contrast, excites the target nucleus to a higher energy level, resulting in some kinetic energy loss to internal excitation. Other direct reaction mechanisms include radiative capture, where the incident particle is absorbed and a gamma ray is emitted, and transfer reactions, such as stripping or pickup, where nucleons are exchanged between projectile and target without forming a compound nucleus. The energetics of any nuclear reaction are quantified by the Q-value, defined as the difference in rest mass energy between initial and final states: Q = (m_{\text{initial}} - m_{\text{final}}) c^2 where m represents atomic masses and c is the speed of light; positive Q-values indicate exoergic (energy-releasing) reactions, driven by the binding energy curve's shape favoring heavier products for light nuclei or lighter fragments for heavy ones. The probability of a reaction occurring is described by its cross-section, typically measured in barns (1 barn = 10^{-28} m²), which quantifies the effective interaction area and depends on incident energy, nuclear structure, and quantum tunneling effects. Nuclear fission is a process where a heavy nucleus splits into two or more lighter fragments, often accompanied by neutron emission and significant energy release of around 200 MeV per event. It can occur spontaneously in certain isotopes, such as , which has a half-life of approximately 8 × 10^{15} years for this decay mode due to quantum tunneling through the fission barrier. Induced fission, more commonly studied, is triggered by the absorption of a neutron or other particle, as discovered by and in 1938 when they observed barium fragments from neutron-bombarded . In the liquid drop model, the fission barrier height for actinides like is typically 5-6 MeV, representing the energy required to deform the nucleus into an unstable configuration before scission. Nuclear fusion combines light nuclei to form a heavier one, releasing energy when the binding energy per nucleon increases, as seen in stellar interiors. A key example is the proton-proton (p-p) chain, dominant in sun-like stars, where four protons fuse stepwise into helium-4, producing positrons, neutrinos, and 26.7 MeV total energy, with the net reaction $4^1\text{H} \rightarrow ^4\text{He} + 2e^+ + 2\nu_e. For fusion to proceed at achievable temperatures, quantum tunneling is essential to penetrate the —the electrostatic repulsion between positively charged nuclei—which for proton-proton fusion has a height of about 1 MeV but effective penetration at stellar core temperatures around 15 million K. Cross-sections for fusion reactions peak at specific energies due to resonances, enabling controlled studies in laboratories. Particle accelerators play a crucial role in probing nuclear reactions by providing high-energy beams to measure cross-sections and simulate astrophysical or terrestrial conditions. Facilities like the Large Hadron Collider (LHC) at CERN accelerate heavy ions, such as lead nuclei, to relativistic speeds for proton-nucleus (pA) and nucleus-nucleus collisions, achieving center-of-mass energies up to 5.02 TeV per nucleon pair and luminosities exceeding 10^{27} cm^{-2} s^{-1} to study quark-gluon plasma formation and nuclear structure effects. These experiments yield precise cross-section data for heavy-ion interactions, informing models of reaction dynamics and energy dissipation.

Isotopes and Nuclear Stability

Isotopes are variants of a chemical element that have the same atomic number Z (number of protons) but different mass numbers A due to varying numbers of neutrons N, where A = Z + N. For instance, carbon-12 (^{12}\mathrm{C}, with Z=6, N=6) is a stable isotope that does not undergo radioactive decay, while carbon-14 (^{14}\mathrm{C}, with Z=6, N=8) is unstable and decays via beta emission to nitrogen-14, with a half-life of approximately 5,730 years. Nuclear stability is determined by the balance between protons and neutrons, visualized in the valley of stability on a chart of nuclides, where stable isotopes cluster around an optimal neutron-to-proton ratio N/Z. For light nuclei (Z < 20), stable isotopes typically have N/Z \approx 1, reflecting near-equal numbers of protons and neutrons to counterbalance Coulomb repulsion with the strong nuclear force. In heavier nuclei, the ratio increases to N/Z \approx 1.5 to provide additional neutrons that enhance stability against proton repulsion without adding more charge. Pairing effects further influence stability, with even-even nuclei (even Z and even N) being the most stable due to complete pairing of nucleons into spin-zero states, minimizing energy; odd-odd nuclei (odd Z and odd N) are the least stable as unpaired nucleons increase excitation energy. This odd-even staggering manifests in half-lives, where even-even isotopes often exhibit longer half-lives compared to neighboring odd-A or odd-odd ones for similar decay modes, a pattern observed across the nuclear chart and attributed to pairing correlations in the shell model. Unstable isotopes decay to approach the valley of stability through specific modes: alpha decay, predominant in heavy nuclei (Z > [82](/page/82)) where nuclei are emitted to reduce size and charge, favored in even-even emitters to preserve ; (minus or plus), which adjusts N/Z imbalances by converting a neutron to a proton (or vice versa) via , common in neutron-rich or proton-rich isotopes; and gamma decay, which de-excites metastable states without changing Z or N, often following alpha or beta decay. The half-life, the time for half of a sample to decay, quantifies instability and varies widely from fractions of seconds to billions of years, influenced by binding energy per nucleon and proximity to the stability valley. In superheavy elements, predictions of an "island of stability" around Z \approx 114-126 and N \approx 184 suggest longer half-lives due to closed shells; for example, oganesson-294 (^{294}\mathrm{Og}, Z=118, N=176) was synthesized in 2002 via calcium-48 bombardment of californium-249, exhibiting a half-life of about 0.7 milliseconds, though isotopes closer to the predicted island may persist longer.

References

  1. [1]
    DOE Explains...Nuclei - Department of Energy
    In 1911, Ernest Rutherford discovered that at the core of every atom is a nucleus. Atomic nuclei consist of electrically positive protons and electrically ...
  2. [2]
    May, 1911: Rutherford and the Discovery of the Atomic Nucleus
    May 1, 2006 · Rutherford carried out a fairly simple calculation to find the size of the nucleus, and found it to be only about 1/100,000 the size of the atom ...Missing: composition forces
  3. [3]
    The Atom and Atomic Structure - Manhattan Project - OSTI.GOV
    The nucleus is made up of protons and neutrons and makes a tiny fraction of the overall size of the atom. The nucleus is only about 10-12 cm across, while the ...
  4. [4]
    Nuclear propertie
    The average radius of a nucleus with A nucleons is R = R0A1/3, where R0 = 1.2*10-15 m. The volume of the nucleus is directly proportional to the total number of ...
  5. [5]
    The Nucleus
    It is found that nuclear radii range from 1-10 ´ 10-15 m. This radius is much smaller than that of the atom, which is typically 10-10 m. Thus, the nucleus ...Missing: diameter | Show results with:diameter
  6. [6]
    DOE Explains...The Strong Force - Department of Energy
    The strong force is the force that holds subatomic particles together to form larger subatomic particles.
  7. [7]
    Science | Inquiring Minds | Questions About Physics - Fermilab
    Apr 28, 2014 · The strong force also acts between protons and neutrons in an atomic nucleus much in the same way that simple chemicals are held together by ...
  8. [8]
    The Strong Nuclear Force
    The strong nuclear force is the strongest of four basic forces, holding subatomic particles together via meson exchange, with a short range.
  9. [9]
    [PDF] Chapter 2 The Atomic Nucleus
    In 1896, A.H. Becquerel discovered penetrating radiation. In 1897, J.J. Thomson showed that electrons have negative electric charge and come from ordinary ...
  10. [10]
    1 Overview | Nuclear Physics: Exploring the Heart of Matter
    Rutherford discovered heavy, apparently pointlike entities at the centers of atoms. He was correct to conclude that nuclei contain most of the mass of an atom, ...
  11. [11]
    Divining the mysteries of the atomic nucleus - C&EN
    Jan 28, 2024 · For decades, scientists pictured this dense nucleus as a roughly spherical ball of protons and neutrons, held together by the strongest force ...
  12. [12]
    Ancient Atomism - Stanford Encyclopedia of Philosophy
    Oct 18, 2022 · Although the Greek term atomos is most commony associated with the solid and impenetrable bodies posited by Leucippus and Democritus, Plato's ...
  13. [13]
    A new system of chemical philosophy : Dalton, John, 1766-1844
    Jan 15, 2008 · 1808-27. Topics: Atomic theory, Chemistry, Inorganic. Publisher ... FULL TEXT download · download 1 file · HOCR download · download 1 file.
  14. [14]
    [PDF] Philosophical Magazine Series 6 XXIV. On the structure of the atom
    J. J. Thomson. To cite this Article Thomson, J. J.(1904) 'XXIV. On the structure of the atom: an investigation of the stability and periods of oscillation of ...
  15. [15]
    [PDF] Sur les radiations émises par phosphorescence. Note de M. HENRI ...
    Sur les radiations invisibles émises par les corpsphosphorescents. Note ... laissant en place les lamelles du sel d'uranium. Le soleil ne s'étant pas.
  16. [16]
    Marie and Pierre Curie and the discovery of polonium and radium
    Dec 1, 1996 · At the end of June 1898, they had a substance that was about 300 times more strongly active than uranium. In the work they published in July ...
  17. [17]
    [PDF] From cathode rays to alpha particles to quantum of action - SciSpace
    Rutherford's response to Geiger was indeed prophetic: “Why not let him see if any alpha particles can be scattered through a large angle?” (reproduced in Wilson ...Missing: expectations | Show results with:expectations
  18. [18]
    On a diffuse reflection of the α-particles - Journals
    A small fraction of the α-particles falling upon a metal plate have their directions changed to such an extent that they emerge again at the side of incidence.Missing: URL | Show results with:URL
  19. [19]
    [PDF] The Laws of Deflexion of α Particles Through Large Angles
    Geiger and E. Marsden, PM, 25, 604. 1913. The Laws of Deflexion of α Particles Through Large. Angles. H. Geiger and E. Marsden. (Received 1913). In a former ...Missing: URL | Show results with:URL
  20. [20]
    [PDF] LXXIX. The scattering of α and β particles by matter and the structure ...
    It was shown that the number N of the electrons within the atom could be deduced from observations of the scattering of electrified particles. The accuracy of ...
  21. [21]
    The existence of a neutron - Journals
    The radiations showed, however, certain peculiarities, and at my request the beryllium radiation was passed into an expansion chamber and several photographs ...
  22. [22]
    None
    Nothing is retrieved...<|separator|>
  23. [23]
    Bakerian Lecture: Nuclear constitution of atoms - Journals
    The conception of the nuclear constitution of atoms arose initially from attempts to account for the scattering of α-particles through large angles.Missing: URL | Show results with:URL
  24. [24]
    The Middle Ages, 300–1500 - Nature
    No readable text found in the HTML.<|separator|>
  25. [25]
    What is an atom ? | Nuclear Regulatory Commission
    The nucleus (or center) of an atom is made up of protons and neutrons. The number of protons in the nucleus, known as the "atomic number," primarily determines ...
  26. [26]
    The Chemical Components of a Cell - Molecular Biology of ... - NCBI
    The number of protons in the atomic nucleus gives the atomic number. An atom of hydrogen has a nucleus composed of a single proton; so hydrogen, with an ...
  27. [27]
    The Radioactive Atom: An Overview | Radiation and Your Health
    Feb 22, 2024 · Electrons surround the nucleus of an atom because of the attraction between their negative charge and the positive charge of the nucleus. A ...
  28. [28]
    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.
  29. [29]
  30. [30]
  31. [31]
    Livechart - IAEA-NDS
    Where Z is the number of protons, N the number of neutrons, and A = Z + N. A letter besides the A number indicates the ENSDF metastable flag, and a decay ...
  32. [32]
    [PDF] Lecture 1: Nuclear Properties - INPP - Ohio University
    ... isotopes (though this term is often used in lieu of nuclides). • Nuclides with the same N but different Z are isotones. • Nuclides with the same A are isobars.
  33. [33]
    Neutrons and Protons in Atomic Nuclei - Nature
    The nucleus of an atom is composed of neutrons and protons. The number of protons equals the atomic number, and the mass number is the sum of neutrons and  ...
  34. [34]
    Modified structure of protons and neutrons in correlated pairs - Nature
    Feb 20, 2019 · The atomic nucleus is made of protons and neutrons (nucleons), which are themselves composed of quarks and gluons.Missing: atomic nucleus composed
  35. [35]
    [PDF] Robert Hofstadter - Nobel Lecture
    Such experiments showed that nuclear radii obeyed an approximate relationship of the type given in Eq. 1.
  36. [36]
    Nuclear charge densities in spherical and deformed nuclei
    May 18, 2021 · We assess the precision of nuclear charge density calculations by studying the behavior of relativistic and center-of-mass motion corrections.
  37. [37]
    Quadrupole deformation in elastic electron scattering
    Feb 10, 2023 · This feature provides valuable information on the nuclear deformation that can be used as a signature of the oblate or prolate character of the nuclear shape.Missing: evidence | Show results with:evidence
  38. [38]
    Insights into nuclear saturation density from parity violating electron ...
    Jul 14, 2020 · The saturation density is closely related to the interior density of a heavy nucleus, such as ^{208}Pb. We use parity violating electron ...
  39. [39]
    Saturation of nuclear matter in the relativistic Brueckner-Hatree-Fock ...
    The saturation of symmetric nuclear matter (SNM), i.e., the energy per nucleon reaches a minimum of about −16 MeV around a density of 0.16 fm − 3 neglecting the ...Letter · 3. Results And Discussions · Declaration Of Competing...
  40. [40]
    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 ...
  41. [41]
    The nuclear charge radius of 13C - PMC - PubMed Central
    Jul 7, 2025 · The size is a key property of a nucleus. Accurate nuclear radii are extracted from elastic electron scattering, laser spectroscopy, ...
  42. [42]
    [PDF] NUCLEAR SCIENCE - Lawrence Berkeley National Laboratory
    Note that the strong force acts between quarks by an exchange of gluons. The residual strong force between two protons can be described by the exchange of a ...<|control11|><|separator|>
  43. [43]
    [PDF] arXiv:1308.0103v1 [nucl-th] 1 Aug 2013
    Aug 1, 2013 · The modern perception of the nuclear force is that it is a residual interaction (similar to the van der Waals force between neutral atoms) of ...
  44. [44]
    Quantum chromodynamics - AIP Publishing
    Interactions among the quarks are mediated by vector force particles called gluons, which themselves expe- rience strong interactions. The nuclear force that ...
  45. [45]
    [PDF] The Standard Model of Particle Physics - Nevis Laboratories
    The strong force, quantum chromody- namics or QCD, is mediated by the exchange of massless gluons (g) between quarks that carry a quantum number called color.
  46. [46]
  47. [47]
    6 Symmetry Tests in Nuclear Physics - The National Academies Press
    The ideas of charge symmetry and charge independence in nuclear physics led to the marvelous concept of isospin. Isospin is an abstract way of describing the ...
  48. [48]
    Understanding the Different Types of Forces in Nature - Empowering ...
    It operates at distances on the order of 1 femtometer (1 fm = 10-15 m), and is approximately 100 times stronger than electromagnetism at these scales. This ...
  49. [49]
    [PDF] NuSTEC Class Notes
    in nuclear forces. 3.6 Pion field of classical sources. In Sec. 2.1 the nonrelativistic scalar meson exchange Yukawa potential is derived using classical ...
  50. [50]
    One-pion exchange potential in a strong magnetic field - arXiv
    Oct 30, 2025 · The Yukawa theory of the pion and the nuclear force laid the foundation for modern particle physics Yukawa:1935xg .
  51. [51]
    One Pion Exchange Potential.
    Let us construct the potential arising from one-pion exchange (OPE), it arises from the tree-level graphs with four external nucleons and an off-shell pion.
  52. [52]
    Exchange Particles - HyperPhysics
    That strong interaction was modeled by Yukawa as involving an exchange of pions, and indeed the pion range calculation was helpful in developing our ...
  53. [53]
    Coulomb Barrier for Nuclear Fusion - HyperPhysics
    VC = x 10^ J = x 10^ eV = keV = MeV. The temperature required to provide this energy as an average thermal energy for each ...
  54. [54]
    phy213 - the physics of stellar interiors - occurrence of fusion reactions
    In classical physics, the only way a nucleus can overcome the Coulomb barrier is for it to increase its kinetic energy. We saw when discussing energy generation ...
  55. [55]
    Coupling Constants for the Fundamental Forces - HyperPhysics
    In attributing a relative strength to the four fundamental forces, it has proved useful to quote the strength in terms of a coupling constant. The coupling ...Missing: nucleus | Show results with:nucleus
  56. [56]
    Experimental Test of Parity Conservation in Beta Decay | Phys. Rev.
    Experimental test of parity conservation in beta decay. CS Wu, E. Ambler, RW Hayward, DD Hoppes, and RP Hudson. Columbia University, New York, New York.
  57. [57]
    Contribution of excited states to stellar weak-interaction rates in odd
    May 5, 2016 · Contribution of excited states to stellar weak-interaction rates in odd- A nuclei ... Electroweak interactions in nuclear physics · Nuclear ...
  58. [58]
    The force is strong - IOP Science
    When two nucleons approach within about. 2 fm of each other they are pulled together. ... ive side (distance greater than 0.8 fm) the strong force curve looks ...
  59. [59]
    The drip line: nuclei on the edge of stability - CERN Courier
    The drip line is the demarcation line between the last bound isotope and its unbound neighbour and each chemical element has a lightest (proton drip line) and ...Missing: ~2.5 fm
  60. [60]
    NUCLEAR HALOS - Annual Reviews
    A neutron halo is basically a threshold effect resulting from the presence of a bound state close to the continuum. The combination of the low neutron ...
  61. [61]
    Measurements of Interaction Cross Sections & Nuclear Radii
    55, 2676 – Published 9 December, 1985. DOI: https://doi.org/10.1103/PhysRevLett ... I. Tanihata et al., Phys. Lett. 160B, 380 (1985); I. Tanihata, Hyperfine ...
  62. [62]
    New Binding Energy for the Two-Neutron Halo of Li-11 - Inspire HEP
    The extended radius of a halo nuclide is very sensitive to the minute binding energy of its valence nucleons. The binding energy of Li11 has been measured ...
  63. [63]
    [1405.2394] $^6$He nucleus in halo effective field theory - arXiv
    May 10, 2014 · The Borromean halo nucleus ^6He exhibits such a separation of scales. In this work an EFT in which the degrees of freedom are the valence ...
  64. [64]
    Breakup of the proton halo nucleus 8B near barrier energies - Nature
    Nov 23, 2022 · The dynamics of a nuclear open quantum system could be revealed in the correlations between the breakup fragments of halo nuclei.
  65. [65]
    [PDF] Breakup of proton halo nucleus 8B at near-barrier energies†
    understanding of the reaction dynamics of halo nuclear systems. The breakup mechanism of the proton halo nuclei is rather complicated. The dynamic Coulomb ...Missing: 6He | Show results with:6He
  66. [66]
    [PDF] Study of the deformed halo nucleus 31Ne with Glauber model ... - OSTI
    include an enhanced reaction cross section, enhanced Coulomb breakup cross section, enhanced nucleon removal cross section, and a peaked momentum ...Missing: implications | Show results with:implications
  67. [67]
    The Mechanism of Nuclear Fission | Phys. Rev.
    On the basis of the liquid drop model of atomic nuclei, an account is given of the mechanism of nuclear fission. In particular, conclusions are drawn ...Missing: original | Show results with:original
  68. [68]
    Zur Theorie der Kernmassen - Zeitschrift für Physik
    Zur Theorie der Kernmassen. Published: July 1935. Volume 96, pages 431–458, (1935); Cite this article. Download PDF · Zeitschrift für Physik. Zur Theorie der ...
  69. [69]
    80 Years of the liquid drop—50 years of the macroscopic ...
    Sep 1, 2013 · The liquid-drop model has its origins in the semiempirical mass model usually credited to von Weizsäcker [1] and Bethe and Bacher [2]. Despite ...
  70. [70]
    On Closed Shells in Nuclei. II | Phys. Rev.
    On Closed Shells in Nuclei. II. Maria Goeppert Mayer. Argonne National Laboratory and Department of Physics, University of Chicago, Chicago, Illinois. PDF ...
  71. [71]
    Maria Goeppert Mayer – Facts - NobelPrize.org
    In 1949 Maria Goeppert Mayer and Hans Jensen developed a model in which nucleons were distributed in shells with different energy levels. The model ...
  72. [72]
    Press release: The 1975 Nobel Prize in Physics - NobelPrize.org
    The very important comparison with experimental data was done in three works, written jointly by Bohr and Mottelson, and published in the years 1952-1953.
  73. [73]
    Collective Motion in the Nucleus | Rev. Mod. Phys.
    Ben R. Mottelson, CERN Theoretical Study Division at the Institute for Theoretical Physics, University of Copenhagen, Copenhagen, Denmark.Missing: paper | Show results with:paper
  74. [74]
    [1703.04640] The many-nucleon theory of nuclear collective ... - arXiv
    Mar 14, 2017 · This paper reviews the microscopic evolution of the collective models and their underlying foundations. In particular, it is shown that the Bohr ...
  75. [75]
    Clusters in nuclei - Scholarpedia
    May 24, 2010 · It was known that conglomerates of nucleons (nuclear clustering) were extremely important in determining the structure of light nuclei.
  76. [76]
    Interacting boson model of collective states I. The vibrational limit
    We propose a unified description of collective nuclear states in terms of a system of interacting bosons. We show that within this model both the vibrational ...Missing: paper | Show results with:paper
  77. [77]
    The Interacting Boson Model | Annual Reviews
    Dec 1, 1981 · The Interacting Boson Model. A Arima and F Iachello; Vol. 31:75-105 (Volume publication date December 1981) ...Missing: 1975 | Show results with:1975
  78. [78]
    Quantum entanglement patterns in the structure of atomic nuclei ...
    Oct 25, 2023 · Yet, nuclear shell-model simulations of mid-mass isotopes indicate that protons and neutrons show very little entanglement [15]. Entanglement is ...
  79. [79]
    Microscopic theory of the nuclear collective model - IOPscience
    This articles reviews the development of a microscopic theory of nuclear collective structure as a submodel of the nuclear-shell model.Missing: paper | Show results with:paper
  80. [80]
    10.3: Nuclear Binding Energy - Physics LibreTexts
    Mar 2, 2025 · Determine the total binding energy (BE) using the equation B ⁢ E = ( Δ ⁢ m ) ⁢ c 2 , where Δ ⁢ m is the mass defect. The binding energy per ...Mass Defect · Example 10 . 3 . 1 : Mass... · Graph of Binding Energy per...
  81. [81]
    Penning-Trap Mass Measurements in Atomic and Nuclear Physics
    Oct 19, 2018 · Penning-trap mass spectrometry in atomic and nuclear physics has become a well-established and reliable tool for the determination of atomic ...
  82. [82]
    Binding energy per nucleon
    NUCLIDE, NUCLIDE MASS IN atomic mass units, MASS minus MASS NUMBER IN atomic mass units ... 2, 26-Iron-56, 55.9349421, -0.06505790, 8.7902483214286, 492.253906 ...Missing: exact | Show results with:exact
  83. [83]
    [PDF] 22.101 Applied Nuclear Physics (Fall 2006)
    The binding energy concept is useful for discussing the calculation of nuclear masses and of energy released or absorbed in nuclear reactions. Binding Energy ...
  84. [84]
    1.2: Binding energy and Semi-empirical mass formula
    Mar 3, 2022 · The Coulomb term seems to indicated that it would be favorable to have less protons in a nucleus and more neutrons. However, this is not the ...
  85. [85]
    Weizsaecker Formula - Semi-empirical Mass Formula - Nuclear Power
    The Weizsaecker formula is a semi-empirical mass formula, a refined liquid drop model for nuclear binding energy, used to derive binding energies and masses.
  86. [86]
    [PDF] Lecture 14: Reactions Overview - INPP - Ohio University
    ... elastic scattering. •The projectile and target aren't transmuted and ... • The Q-value is equal to any excess kinetic energy resulting from the reaction,.
  87. [87]
    [PDF] RDCH 702: Lecture 3, Nuclear Reactions - UNLV Radiochemistry
    → Need more energy than Q value for reaction to occur. * Reaction ... ▫ elastic and inelastic scattering. ▫ compound-nucleus formation,. ▫ direct ...
  88. [88]
    Nuclear reaction - Book chapter - IOPscience
    We discuss elastic and inelastic scattering. Q value of nuclear reaction is defined. Two-body non-relativistic Q value equation is solved and expression of ...<|separator|>
  89. [89]
    SPONTANEOUS FISSION OF NUCLEI - IOP Science
    In December, 1938, there appeared the first report of the discovery by Hahn and Strassmann of the fission of uranium nuclei by neutrons.1. Further reports of ...
  90. [90]
    [PDF] Nuclear Fission - OSTI.GOV
    Oct 19, 2012 · In December 1938, Hahn and Strassmann sent a manuscript to Naturwissenschaften reporting they had detected the element barium after ...
  91. [91]
    [PDF] Fission Cross Sections_FIESTA17-Rev12 - LANL
    1938: OYo Hahn and Fritz Strassmann irradiate ... nuclear fission based on the liquid drop model ... • Improved confidence in our current knowledge of neutron-‐ ...
  92. [92]
    32.5 Fusion – College Physics - UCF Pressbooks
    Nuclear fusion is a reaction in which two nuclei are combined to form a larger nucleus. · Fusion is the source of energy in stars, with the proton-proton cycle,.Missing: chain | Show results with:chain
  93. [93]
    Nuclear Fusion - HyperPhysics
    The reaction yields 17.6 MeV of energy but to achieve fusion one must penetrate the coulomb barrier with the aid of tunneling, requiring very high temperatures ...Missing: chain | Show results with:chain
  94. [94]
    [PDF] Physics with high-luminosity proton-nucleus collisions at the LHC
    Aug 10, 2025 · Compared to AA collisions, pA collisions offer higher nucleon–nucleon CM energies, higher luminosity, and a cleaner environment, free from the ...
  95. [95]
    Opportunities for new physics searches with heavy ions at colliders
    Dedicated runs of p-O [2] and light-ion collisions at the LHC are required to improve the modeling and tuning of all nuclear effects in the current hadronic MC ...
  96. [96]
    11.3: Stable and Unstable Isotopes - Chemistry LibreTexts
    Dec 20, 2023 · Carbon-12, with six protons and six neutrons, is a stable nucleus, meaning that it does not spontaneously emit radioactivity. Carbon-14, with ...Missing: definition | Show results with:definition
  97. [97]
    Isotope - Nuclear Regulatory Commission
    For example, carbon-12 and carbon-13 are stable, but carbon-14 is unstable and radioactive.Missing: definition | Show results with:definition
  98. [98]
    Nuclear Magic Numbers - Chemistry LibreTexts
    Jan 29, 2023 · The two main factors that determine nuclear stability are the neutron/proton ratio and the total number of nucleons in the nucleus. Introduction.
  99. [99]
    19.1 Nuclear Structure and Stability – Chemistry Fundamentals
    The others are the electromagnetic force, the gravitational force, and the nuclear weak force.) This force acts between protons, between neutrons, and between ...<|control11|><|separator|>
  100. [100]
    13.2 Nuclear stability and the chart of nuclides - Fiveable
    Position-based stability assessment. Nuclei near valley of stability center generally more stable; N/Z ratio of stable nuclei increases with atomic number.
  101. [101]
    Stable Nuclides - an overview | ScienceDirect Topics
    It shows that an even-even combination of Z and N gives high stability and odd-odd the least. Table 1.2. Stable nuclides.
  102. [102]
    Measurement and microscopic description of odd–even staggering ...
    Apr 13, 2020 · This odd–even staggering, ubiquitous throughout the nuclear landscape, varies with the number of protons and neutrons, and poses a substantial ...Missing: half- | Show results with:half-
  103. [103]
    Favored 𝛼 -decay half-lives of odd- A and odd-odd nuclei using an ...
    Aug 8, 2022 · Meanwhile, the theoretical half-lives agree very well with the experimental data for the favored decays of the odd- and odd-odd nuclei with a ...<|control11|><|separator|>
  104. [104]
    20.1: Nuclear Stability and Radioactive Decay - Chemistry LibreTexts
    Apr 27, 2019 · The most common are alpha and beta decay and gamma emission, but the others are essential to an understanding of nuclear decay reactions.Missing: imbalance | Show results with:imbalance
  105. [105]
    Modes of Radioactive Decay | Radiology Key
    Apr 3, 2016 · The most important modes of radioactive decay are: alpha, beta minus, beta plus, electron capture, gamma and internal conversion.
  106. [106]
    Radioactive Decay - an overview | ScienceDirect Topics
    The transition often is to an excited nuclear state of the daughter which decays by emission of one or more gamma photons. Beta decay (β−): P Z A → D Z + ...<|control11|><|separator|>
  107. [107]
    Systematic study of $α$ decay for odd-$A$ nuclei within a ... - arXiv
    Mar 18, 2019 · We found that the odd-even staggering effect may play a more important role on spontaneous fission than \alpha decay. The calculated half-lives ...
  108. [108]
    Colloquium: Superheavy elements: Oganesson and beyond
    Jan 22, 2019 · The nucleus Og 294 , produced in Dubna ( Oganessian et al., 2006 ) with a ∼ 0.5 p b cross section, marks the current limit of nuclear charge and ...
  109. [109]
    Investigation of probabilities of formation and decay of superheavy ...
    Aug 12, 2021 · The heaviest element known today is Oganesson (294Og) with Z = 118 produced in the complete fusion reaction 48Ca + 249Cf. The production of ...<|control11|><|separator|>