Fact-checked by Grok 2 weeks ago

Quantum number

In , quantum numbers are discrete integers or half-integers that arise as eigenvalues of specific operators in the solutions to the , characterizing the unique quantum states of particles such as electrons in atoms and defining properties like , , and orientation. These numbers result from boundary conditions imposed on the wave function, ensuring it remains finite and single-valued within the system's potential, thus quantizing quantities. For electrons in multi-electron atoms, four quantum numbers fully specify the state according to the Pauli exclusion principle, which states that no two electrons can share the same set of all four values. The principal quantum number (n) is a positive integer (1, 2, 3, ...) that primarily determines the electron's energy level and the average size of its orbital, with higher n corresponding to higher energy and larger orbitals. The azimuthal quantum number (l), ranging from 0 to n-1, describes the orbital's shape (e.g., l=0 for s orbitals, l=1 for p, l=2 for d), reflecting the magnitude of the orbital angular momentum. The magnetic quantum number (m_l), taking integer values from -l to +l, specifies the orbital's orientation relative to an external magnetic field. Finally, the spin quantum number (m_s) is ±1/2, accounting for the electron's intrinsic spin angular momentum, a relativistic effect incorporated into the Dirac equation but approximated in non-relativistic treatments. Quantum numbers originated in early models like Bohr's 1913 atomic theory, which introduced the principal number n to quantize electron orbits, and evolved through Sommerfeld's 1916 extension adding l for elliptical paths, culminating in the full quantum mechanical framework by Schrödinger and Heisenberg in 1926. Beyond atomic electrons, quantum numbers describe states in nuclear physics (e.g., proton and neutron shells), molecular orbitals, and even exotic systems like quarks in quantum chromodynamics, underscoring their foundational role in modern physics. The allowed combinations of these numbers dictate electron configurations, chemical bonding, and periodic trends, forming the basis for understanding atomic spectra and material properties.

Fundamentals

Definition and Classification

Quantum numbers are discrete values that characterize the of a , arising from the solutions to the or such as the . These values emerge as parameters that label the eigenfunctions and eigenvalues of the time-independent for bound systems, ensuring that physical observables like energy take on quantized, non-continuous forms. In quantum mechanics, quantum numbers specify key properties of a system's state, including energy levels, angular momenta, and other measurable observables./01%3A_Summary_of_things_you_should_already_know/1.09%3A_Good_quantum_numbers) They arise from a complete set of commuting observables, whose simultaneous eigenstates provide a unique description of the system; incompatible observables, whose operators do not commute, cannot share a common set of definite quantum numbers. This framework ensures that measurements of compatible observables yield precise, reproducible values, while transitions between states obey selection rules dictated by the matrix elements of interaction operators. Quantum numbers can be classified in several ways based on their physical origin and behavior in composite systems. Additive quantum numbers, such as total components, combine vectorially or by summation for multi-particle states, reflecting continuous symmetries in the . In contrast, multiplicative quantum numbers, like or charge conjugation, take values such as ±1 and multiply for composite systems, arising from symmetries. Additionally, they are distinguished as orbital types (describing spatial wavefunction properties) versus types (characterizing intrinsic particle ), and as primary quantum numbers (fundamental labels from single-particle solutions) versus secondary ones (derived from , such as total angular momentum j). A representative example is the set of four quantum numbers for electrons in atoms: the principal quantum number n (determining the energy shell, n = 1, 2, [3, \dots](/page/3_Dots)), the l (specifying orbital , l = 0, [1](/page/1), \dots, n-1), the m_l (orienting the orbital, m_l = -l, \dots, +l), and the m_s (indicating spin projection, m_s = \pm 1/2). These form a complete set for specifying individual states, enforcing the that no two electrons share the same set.

General Properties and Conservation

Quantum numbers are eigenvalues of Hermitian operators representing measurable observables in quantum mechanics. A key property is their compatibility: quantum numbers are compatible if their corresponding operators commute, enabling the system to possess simultaneous eigenstates for those observables. For instance, the operators for the square of the orbital angular momentum \hat{L}^2 and its z-component \hat{L}_z commute, [ \hat{L}^2, \hat{L}_z ] = 0, allowing states to be labeled by both the quantum numbers l (for \hat{L}^2 | l, m_l \rangle = \hbar^2 l(l+1) | l, m_l \rangle) and m_l (for \hat{L}_z | l, m_l \rangle = \hbar m_l | l, m_l \rangle). This compatibility ensures that measurements of compatible observables do not disturb each other, forming a complete set of commuting observables (CSCO) that uniquely specifies the state. In isolated quantum systems, many quantum numbers are conserved quantities, arising from symmetries via , which establishes a one-to-one correspondence between continuous symmetries of the action and conserved currents. For example, time-translation invariance (a symmetry of the ) conserves , labeled by the principal quantum number in bound systems, while spatial rotational invariance conserves total , preserving quantum numbers like total j. Discrete symmetries, such as under spatial inversion, also lead to conserved multiplicative quantum numbers in systems invariant under those transformations. These conservation laws hold exactly for non-interacting or symmetrically invariant Hamiltonians but can be approximate otherwise. For multi-particle systems, quantum numbers exhibit additivity or multiplicativity depending on the . Additive quantum numbers, such as total spin angular momentum, combine vectorially as the sum over individual particle : \mathbf{S} = \sum_i \mathbf{s}_i, yielding a total S that labels the composite state's magnitude. In contrast, multiplicative quantum numbers like combine as the product of individual parities: P = \prod_i p_i, where each p_i = \pm 1, resulting in an overall parity eigenvalue for the system. These properties facilitate the of composite states in , , and . Quantum transitions, such as those induced by electromagnetic interactions, are governed by selection rules that restrict allowable changes in quantum numbers, determined by the matrix elements of the interaction operator between initial and final states. For electric (E1) transitions, the dominant radiative process in atoms, the selection rule is \Delta l = \pm 1 for the orbital angular momentum quantum number, alongside \Delta m_l = 0, \pm 1 and parity change, ensuring non-zero transition probability only for allowed \Delta l. These rules arise from the vector nature of the dipole operator and symmetry considerations. Despite their utility, quantum numbers have limitations: not every qualifies as a quantum number, particularly if its does not commute with the or other relevant symmetries, preventing conserved or simultaneously measurable eigenvalues. In interacting systems, such as those with perturbations that break exact symmetries, "good" quantum numbers become approximate, as states mix and eigenstates deviate from ideal simultaneous eigenvectors, requiring for corrections. For example, spin-orbit coupling in atoms renders individual l and s approximate labels, with total j becoming the good quantum number instead.

Historical Development

Origins in Atomic Spectra

The observation of discrete spectral lines in atomic spectra provided the first empirical evidence necessitating quantum numbers to describe atomic energy levels. In 1885, Johann Balmer identified a series of visible lines that followed an relating wavelengths to integers, later generalized by in 1889 as \frac{1}{\lambda} = R \left( \frac{1}{n_1^2} - \frac{1}{n_2^2} \right), where R is the and n_1, n_2 are integers with n_2 > n_1. This formula implied quantized energy levels in the , as continuous orbits would produce a continuum of wavelengths rather than discrete lines. Niels Bohr's 1913 model of the introduced the principal quantum number n to explain these discrete levels through quantized . Bohr postulated that the electron's orbital is L = n \hbar, where n = 1, 2, [3, \dots](/page/3_Dots) and \hbar = h / 2\pi with h Planck's , leading to stationary orbits with energies E_n = -\frac{13.6}{n^2} eV. This quantization rule reproduced the exactly, marking the first systematic use of a quantum number in . In 1896, discovered the splitting of spectral lines in a , known as the , which required an additional quantum number to account for the observed multiplet structure. Arnold Sommerfeld's 1916 extension of the incorporated relativistic corrections and elliptical orbits, introducing the k (later denoted l, with l = 1, 2, \dots, n) to describe the eccentricity of orbits and explain splittings in spectra. To interpret the within this framework, Sommerfeld also proposed the m_l (with m_l = -l, \dots, +l), representing the projection of orbital along the field direction, thus enabling the prediction of line splittings proportional to m_l. The discovery of electron spin added further quantum numbers in 1925, when and proposed that the possesses an intrinsic with s = 1/2, introducing the m_s = \pm 1/2 to account for doublet splittings in alkali metal spectra. This hypothesis was supported by the 1922 Stern-Gerlach experiment, which demonstrated the deflection of silver atoms into two beams in an inhomogeneous , confirming the quantized spin projection. Despite these advances, the relied on ad hoc quantization rules, such as action integrals being integer multiples of h, which successfully described some phenomena but failed for systems like the specific heats of polyatomic gases or intensities of lines. These limitations, including the inability to derive transition probabilities systematically, highlighted the need for a more fundamental framework, paving the way for the development of in 1925–1926 through by , , and , and wave mechanics by .

Expansion to Nuclear and Particle Physics

The discovery of the neutron by in 1932 marked a pivotal shift in understanding nuclear structure, necessitating the extension of quantum numbers to describe uncharged particles within the alongside protons. This revelation implied that atomic quantum numbers, originally developed for electrons, required adaptation for nucleons, as protons and neutrons occupy distinct quantum states despite their similar masses, influencing nuclear binding and stability. Shortly thereafter, introduced the concept of in 1932 to account for the near-symmetry between protons and neutrons in interactions, treating them as two states of a single isotopic with isospin quantum number I = 1/2. This formalism extended angular momentum-like quantum numbers to the charge of , facilitating the description of strong forces without distinguishing charge at short ranges. Building on this, proposed in 1935 that the strong force is mediated by a , later identified as the ; Yukawa's theory, which assumed the conservation of —a multiplicative quantum number introduced by in 1927 and conserved in strong interactions—further broadened the quantum number framework to meson exchanges. In the 1940s, neutron diffraction experiments, pioneered by Ernest O. Wollan and Clifford G. Shull using reactor-produced beams, confirmed the and quantum numbers of nuclei by revealing magnetic patterns consistent with spins of $1/2. These observations validated the application of orbital angular momentum l and total angular momentum j to nuclear constituents, analogous to electrons, and underscored the role of neutron-proton in nuclear spectra. The , independently developed by and J. Hans D. Jensen in 1949, formalized this expansion by assigning quantum numbers such as nuclear orbital angular momentum l_n and spin to protons and neutrons in filled shells, explaining and nuclear stability through Pauli exclusion principles. This model drew direct parallels to atomic shell structures but incorporated strong spin-orbit coupling for nucleons, predicting ground-state properties for a wide range of isotopes. Advancing into , and proposed the in 1964, assigning fractional , spin $1/2, and quantum numbers (up, down, strange) to fundamental constituents of hadrons, thereby generalizing quantum numbers to subnuclear scales. This framework resolved the proliferation of particles observed in accelerators by composing baryons and mesons from combinations, with and emerging as composite symmetries. However, the strong interactions governing quarks posed challenges, as perturbative methods failed at low energies; the development of (QCD) in the early 1970s, with demonstrated in 1973 by , , and David Politzer, incorporated —first proposed by Oscar W. Greenberg in 1964—as a new SU(3) quantum number with three types (red, green, blue), ensuring color neutrality in hadrons and enabling at high energies. This non-Abelian revolutionized particle descriptions, confining quarks within color singlets while allowing quantum numbers like and to remain conserved.

Quantum Numbers in Atomic Systems

Hydrogen-like Atoms

In hydrogen-like atoms, consisting of a nucleus with atomic number Z and a single electron, the quantum numbers arise from solving the time-independent Schrödinger equation for the electron in the Coulomb potential V(r) = -\frac{Z e^2}{4\pi \epsilon_0 r}. The equation is -\frac{\hbar^2}{2\mu} \nabla^2 \psi(\mathbf{r}) + V(r) \psi(\mathbf{r}) = E \psi(\mathbf{r}), where \mu is the reduced mass of the electron-nucleus system and E is the energy eigenvalue. Due to the spherical symmetry of the potential, the equation separates in spherical coordinates (r, \theta, \phi) into radial and angular parts, with the wave function expressed as \psi_{nlm_l}(r, \theta, \phi) = R_{nl}(r) Y_{l m_l}(\theta, \phi), where Y_{l m_l} are spherical harmonics. This separation yields the four quantum numbers that fully specify the electron's state in the non-relativistic approximation. The principal quantum number n emerges from the radial and takes positive values n = 1, 2, [3, \dots](/page/3_Dots). It determines the levels as E_n = -\frac{\mu Z^2 e^4}{8 \epsilon_0^2 h^2 n^2} = -\frac{13.6 \, \mathrm{[eV](/page/EV)} \cdot Z^2}{n^2}, independent of other quantum numbers, leading to degeneracy in the . The radial R_{nl}(r) has n - l - 1 nodes, reflecting the oscillatory behavior confined by the centrifugal barrier and attraction. The l, also known as the orbital quantum number, arises from both the radial and angular equations and ranges from $0 to n-1. It characterizes the magnitude of the as \mathbf{L}^2 = l(l+1) \hbar^2, with eigenvalues derived from the in the angular solution. Conventionally, l = 0, 1, 2, 3, \dots correspond to s, p, d, f subshells, respectively, influencing the spatial extent and shape of the orbitals. The m_l specifies the z-component of the and takes integer values from -l to +l in steps of 1. It determines L_z = m_l [\hbar](/page/H-bar) from the \phi-dependent part of the , which are eigenfunctions of the rotation operator around the z-axis. In the absence of an external , states with different m_l but the same n and l are degenerate, contributing to the (2l + 1)-fold orbital degeneracy. The m_s accounts for the 's intrinsic , introduced to explain in atomic spectra. The has s = 1/2, so m_s = \pm 1/2, with \mathbf{S}^2 = s(s+1) [\hbar](/page/H-bar)^2 = (3/4) [\hbar](/page/H-bar)^2 and S_z = m_s [\hbar](/page/H-bar). This doubles the degeneracy of each orbital state, as is independent of the orbital motion in the non-relativistic limit. The combination of these quantum numbers implies the , which states that no two can occupy the same state defined by the set \{n, l, m_l, m_s\}, ensuring unique labeling of states in multi-electron systems. For a given principal quantum number n, the total number of states in the shell is $2n^2, arising from n possible values of l, (2l + 1) for m_l, and 2 for m_s. This complete specification \{n, l, m_l, m_s\} uniquely identifies each allowed electron state in hydrogen-like atoms.

Multi-electron Atoms and Shells

In multi-electron atoms, the arrangement of electrons is governed by the Pauli exclusion principle, which requires that no two electrons share the same set of four quantum numbers (n, l, m_l, m_s). This leads to the formation of electron shells and subshells, where electrons occupy orbitals in a manner that minimizes the total energy while respecting quantum mechanical constraints. The principal quantum number n defines the shell, with the innermost shell designated as K (n=1), followed by L (n=2), M (n=3), N (n=4), and so on. Within each shell, subshells are characterized by the azimuthal quantum number l, ranging from 0 to n-1, and labeled as s (l=0), p (l=1), d (l=2), and f (l=3). The maximum capacity of a subshell is given by 2(2l + 1) electrons, accounting for the 2l + 1 possible m_l values and two possible m_s values (±1/2) per orbital. The dictates that electrons fill orbitals starting from the lowest energy levels, determined primarily by the principal quantum number n and the l through the n + l rule: orbitals with lower n + l values are filled first, and for equal n + l, lower n takes precedence. This results in the filling order 1s < 2s < 2p < 3s < 3p < 4s < 3d < 4p < 5s < 4d < 5p < 6s < 4f < 5d < 6p, and so forth. Due to electron-electron interactions, the energy ordering deviates from the simple hydrogenic case, with penetration effects causing 4s orbitals to be lower in energy than 3d for elements like potassium. For degenerate orbitals within a subshell, Hund's rules determine the ground-state configuration by maximizing the total spin angular momentum S (sum of individual s_i) to achieve the highest multiplicity (2S + 1), thereby minimizing electron repulsion through parallel spins. If multiple states have the same multiplicity, the one with the maximum total orbital angular momentum L (sum of individual l_i) is lowest in energy, as this maximizes the average distance between electrons. Half-filled or fully filled subshells, such as a p^3 or d^5 configuration, exhibit enhanced stability due to these rules, as seen in the ground state of nitrogen (1s^2 2s^2 2p^3). The total angular momentum quantum numbers L and S couple to form the total angular momentum J via LS (Russell-Saunders) coupling, appropriate for light atoms where spin-orbit interaction is weak. Atomic states are denoted by term symbols ^{2S+1}L_J, where L is represented by S for 0, P for 1, D for 2, F for 3, etc., and J ranges from |L - S| to L + S in integer steps. For example, the ground state of carbon is ^3P_0, arising from the 2p^2 configuration with L=1 (P) and S=1 (triplet). These quantum number assignments directly underpin the structure of the , where blocks correspond to the filling of specific subshells: s-block elements (groups 1-2) have ns^1 or ns^2 configurations, p-block (groups 13-18) fill np^1 to np^6, d-block (transition metals, groups 3-12) fill (n-1)d^1 to (n-1)d^{10}, and f-block (lanthanides and actinides) fill (n-2)f. This organization explains periodic trends in chemical properties, such as valence electron count influencing reactivity and ionization energies. Exceptions to the Aufbau order occur when alternative configurations provide greater stability, particularly for half-filled or fully filled subshells. Chromium ([Ar] 4s^1 3d^5) and copper ([Ar] 4s^1 3d^{10}) adopt these arrangements instead of the expected [Ar] 4s^2 3d^4 and [Ar] 4s^2 3d^9, respectively, due to the lowered energy from symmetric electron distribution and reduced electron-electron repulsion in the d subshell.

Nuclear Quantum Numbers

Angular Momentum and Parity

In nuclear physics, the total angular momentum quantum number I characterizes the spin of an atomic nucleus, representing the magnitude of its total angular momentum vector, with possible values ranging from 0 to A/2, where A is the mass number (total number of nucleons). The projection of this angular momentum along a specified axis is given by the magnetic quantum number m_I, taking values from -I to +I in steps of 1./19%3A_Nuclear_Magnetic_Resonance_Spectroscopy/19.01%3A_Theory_of_Nuclear_Magnetic_Resonance) For even-even nuclei (even proton and neutron numbers), the ground state typically has I = 0, while odd-A nuclei exhibit half-integer spins due to the unpaired nucleon. The total nuclear angular momentum \mathbf{J} arises from the vector sum of the orbital angular momentum \mathbf{L} and the total spin angular momentum \mathbf{S}, such that \mathbf{J} = \mathbf{L} + \mathbf{S}./04%3A_Nuclear_Models/4.01%3A_Nuclear_Shell_Model) In the nuclear shell model, individual nucleons occupy orbitals characterized by their orbital angular momentum quantum number l_n (an integer from 0 to some maximum), and their intrinsic spin s = 1/2, coupling to a single-particle total angular momentum j = l_n \pm 1/2; these are then coupled across all nucleons to yield the overall J./04%3A_Nuclear_Models/4.01%3A_Nuclear_Shell_Model) The value of I (or J for excited states) is often inferred from nuclear magnetic moments, which probe the expectation value of the magnetic dipole operator and reveal the relative contributions of orbital and spin components. The parity quantum number \pi describes the behavior of the nuclear wave function under spatial inversion (\mathbf{r} \to -\mathbf{r}), with eigenvalues +1 (even parity) or -1 (odd parity). For a nucleus, the total parity is the product of the intrinsic parities of the constituent nucleons (each +1 for protons and neutrons) and the orbital parity factor (-1)^{\sum l_n}, where the sum is over all occupied single-particle orbitals. In the shell model, even-A nuclei in closed-shell configurations frequently exhibit even parity (\pi = +), as the paired nucleons fill orbitals with even total \sum l_n./04%3A_Nuclear_Models/4.01%3A_Nuclear_Shell_Model) In nuclear transitions, such as gamma decay, conservation of angular momentum imposes selection rules on the change in total angular momentum \Delta I = 0, \pm 1 (with the restriction that $0 \to 0 transitions are forbidden), where the photon carries away angular momentum of multipolarity L \geq |\Delta I|./07%3A_Radioactive_Decay_Part_II/7.01%3A_Gamma_Decay) The parity change \Delta \pi further determines the transition type: no parity change allows electric $2^L-pole (EL) or magnetic $2^L-pole (ML) radiation, while a parity change permits EL for odd L or ML for even L, influencing the decay rate and multipolarity./07%3A_Radioactive_Decay_Part_II/7.01%3A_Gamma_Decay) Representative examples illustrate these quantum numbers in the shell model framework. The deuteron (^2\mathrm{H}), consisting of a proton and neutron in a predominantly l = 0 orbital with total spin 1, has ground-state quantum numbers I = 1, \pi = +./05%3A_Nuclear_Structure/5.02%3A_The_Deuteron) Similarly, the ground state of ^{12}\mathrm{C}, a doubly magic nucleus with closed p-shells for both protons and neutrons, is assigned I = 0, \pi = +, consistent with paired nucleons and even parity orbitals. These quantum numbers are measured experimentally through techniques such as nuclear magnetic resonance (NMR), which detects resonant transitions between m_I states in a magnetic field to determine I and the magnetic moment, and beta decay spectroscopy, where angular correlations in emitted particles reveal spin and parity via selection rules and asymmetry patterns.

Isospin and Baryon Number

The baryon number B is an additive quantum number that labels the number of baryons in a system, assigned as B = 1 for baryons such as protons and neutrons, and B = \frac{1}{3} for each , the fundamental constituents of baryons. This quantum number is strictly conserved in all strong and electroweak interactions within the , with no observed violations in experiments probing energies up to the electroweak scale. Isospin I emerges from an approximate SU(2) symmetry of the strong nuclear force, which treats protons and neutrons not as distinct particles but as the two components of an isospin doublet for the nucleon, with total isospin I = \frac{1}{2}, third component I_3 = +\frac{1}{2} for the proton, and I_3 = -\frac{1}{2} for the neutron. The total isospin of a nucleus is obtained by vector addition of the individual nucleons' isospins; for instance, ground states of even-even nuclei (with even numbers of protons and even numbers of neutrons) typically have total I = 0 due to pairing effects. The third component I_3 for the entire nucleus is the algebraic sum of the individual I_3 values, which connects to the total electric charge Q through the Gell-Mann–Nishijima formula: Q = I_3 + \frac{B + S}{2} where S is the strangeness quantum number (zero for ordinary nuclear matter without strange quarks). This isospin formalism underpins key applications in nuclear structure, such as the symmetry observed in mirror nuclei—pairs like ^{14}\mathrm{C} (6 protons, 8 neutrons) and ^{14}\mathrm{O} (8 protons, 6 neutrons)—whose binding energies and spectra are nearly identical, demonstrating the charge independence of the nuclear force under strong interactions. However, isospin symmetry is violated by the weak interaction, which distinguishes between protons and neutrons; for example, in beta decay processes, the change \Delta I_3 = \pm 1 accompanies the transformation of a neutron to a proton (or vice versa), breaking the symmetry explicitly. In the modern quark model context, isospin symmetry extends naturally to the light quarks, with the up quark (I_3 = +\frac{1}{2}) and down quark (I_3 = -\frac{1}{2}) forming an I = \frac{1}{2} doublet, mirroring the nucleon structure and reinforcing the underlying SU(2) flavor symmetry of for these flavors.

Quantum Numbers in Particle Physics

Elementary Particles and Spin

In the Standard Model of particle physics, elementary particles are classified by their intrinsic quantum numbers, with spin being a fundamental property that distinguishes fermions from bosons. Fermions, which include quarks and leptons, possess half-integer spin values, such as s = 1/2, and obey the Pauli exclusion principle due to the spin-statistics theorem. This theorem, established by Wolfgang Pauli, dictates that particles with half-integer spin must have antisymmetric wave functions under particle exchange, leading to Fermi-Dirac statistics, while integer-spin particles follow symmetric Bose-Einstein statistics. For example, all six quark flavors (up, down, strange, charm, bottom, top) and the six leptons (electron, muon, tau, and their neutrinos) have spin s = 1/2. The spin quantum number s determines the possible projections along a quantization axis, denoted by the magnetic quantum number m_s, which ranges from -s to +s in integer steps. For spin-1/2 particles like the electron, m_s = \pm 1/2. Massless particles, such as photons (spin s = 1) and originally assumed massless neutrinos, are characterized by helicity \lambda, the projection of spin along the direction of motion, taking values \lambda = \pm s. In the Standard Model, neutrinos are left-handed, meaning only the \lambda = -1/2 state interacts via the weak force, a consequence of the chiral structure of weak interactions. Gauge bosons, which mediate the fundamental forces, include photons (s = 1, electromagnetic), gluons (s = 1, strong), and W/Z bosons (s = 1, weak), all with integer spin and bosonic statistics. The Higgs boson, a scalar particle with s = 0, breaks electroweak symmetry and imparts mass to other particles. Leptons carry distinct lepton numbers as additive quantum numbers conserved in Standard Model interactions: electron lepton number L_e = +1 for the electron and -1 for the positron, with analogous L_\mu and L_\tau for muon and tau families, including their neutrinos. These are violated only in processes beyond the Standard Model, such as neutrinoless double beta decay, which would indicate lepton number non-conservation by two units and Majorana neutrino nature. For quarks, flavor quantum numbers label the six types across three generations: first (up u, down d), second (charm c, strange s), third (top t, bottom b). The third component of isospin I_3 assigns +1/2 to up-type quarks (u, c, t) and -1/2 to down-type (d, s, b), while strangeness S = -1 for the strange quark, charm C = +1 for charm, and similarly for bottomness B = -1 and topness T = +1. These flavors mix via the Cabibbo-Kobayashi-Maskawa matrix in weak interactions, but are otherwise conserved in strong and electromagnetic processes. Representative examples illustrate these properties: the electron has s = 1/2, m_s = \pm 1/2, and L_e = 1; the up quark has s = 1/2, I_3 = +1/2, and belongs to the first generation. The photon exemplifies a spin-1 boson with \lambda = \pm 1 (transverse polarizations), while the Higgs boson, discovered at the LHC, has confirmed spin-0 through angular correlations in decay products like H \to ZZ \to 4\ell. Experimental verification of these quantum numbers relies on high-energy colliders like the Large Hadron Collider (LHC), where ATLAS and CMS detectors measure spin via decay angular distributions and production kinematics. For instance, the Higgs spin was determined to be 0 (with spin-2 excluded at >99% confidence) from analyses of its decays to photons, W/Z bosons, and taus. Similarly, and spins are inferred from jet substructure and lepton polarizations in events like decays, confirming the assignments.

Multiplicative and Additive Quantum Numbers

In , additive quantum numbers are conserved quantities that add linearly across initial and final states in interactions, reflecting underlying symmetries in the . Key examples include the B, which counts the number of valence quarks minus antiquarks (with quarks having B = +1/3), the total L, and the Q. These are preserved in strong, electromagnetic, and weak interactions at the perturbative level, though non-perturbative electroweak processes, such as transitions, can violate B and L while conserving B - L. Quarks carry specific additive quantum numbers, such as up-type quarks with Q = +2/3 and down-type with Q = -1/3, ensuring overall charge neutrality in hadrons. Multiplicative quantum numbers, by contrast, arise from discrete symmetries and assign phase factors of \pm 1 to particle states under transformation, multiplying across states rather than adding. Prominent instances are charge conjugation C, which interchanges particles and antiparticles; parity P, which reflects spatial coordinates; and time reversal T, which reverses time direction. The CPT theorem guarantees that the combined CPT transformation is a symmetry of all local, Lorentz-invariant quantum field theories with unitary evolution and finite-dimensional spin representations, implying that particles and antiparticles have identical masses, lifetimes, and decay rates. This theorem, first rigorously proven in the context of quantum field theory, underpins the equality of matter and antimatter properties observed experimentally. The combined operator CP is conserved in strong and electromagnetic interactions but violated in weak processes, as evidenced by the 1964 observation of K_L^0 \to \pi^+ \pi^- decays, which should be forbidden under CP invariance for the long-lived neutral kaon. This discovery by Christenson, Cronin, Fitch, and Turlay demonstrated a small (\sim 10^{-3}) admixture of CP-even states in the nominally CP-odd K_L^0, confirming CP violation and necessitating a non-zero phase in the Cabibbo-Kobayashi-Maskawa matrix. Other multiplicative quantum numbers include G-parity, defined as G = C e^{i \pi I_2} where I_2 is the third component of isospin, applicable to isospin multiplets like the pion triplet (G = -1); it is conserved in strong interactions and helps classify hadron states. In supersymmetric extensions of the Standard Model, R-parity is a proposed multiplicative quantum number given by R = (-1)^{3(B - L) + 2s}, where s is spin, introduced in the 1980s to suppress rapid proton decay by forbidding superpartner interactions; its conservation remains unverified, with ongoing searches for violation at colliders. Conservation patterns differ across fundamental interactions, dictating allowed processes. Strong interactions preserve all additive quantum numbers (B, L, Q) and multiplicative ones (C, P, T, G), enforcing strict selection rules for hadron decays and scatterings. Electromagnetic interactions conserve B, L, Q, C, and P but may violate T in principle (though unobserved beyond CP effects via CPT), as they couple to charged states without flavor change. Weak interactions uphold B, L, and Q but violate P (maximal in charged currents), C, and CP (via phase-dependent mixing), while invariably conserving CPT; this leads to processes like where mirror-image asymmetries appear. These rules extend to combined symmetries, with strong and electromagnetic forces respecting CP fully, unlike the weak force. Such quantum numbers impose selection rules that govern decay modes and branching ratios, providing stringent tests of theories. For instance, the electromagnetic decay \pi^0 \to \gamma \gamma proceeds because the neutral pion has C = +1 and the two-photon final state also carries C = (+1) \times (+1) = +1, while odd-photon modes like \pi^0 \to \gamma are forbidden by C conservation; this two-photon channel dominates the \pi^0 lifetime of about $8.5 \times 10^{-17} s. Similarly, G-parity forbids certain pion multiplicities in strong decays, and R-parity would prohibit single superpartner production at colliders if conserved. Violations or conservations thus probe beyond-Standard-Model physics, as in unconfirmed R-parity breaking or precise CP measurements in B mesons.

Symmetries and Advanced Applications

Connection to Symmetry Groups

Quantum numbers arise fundamentally from the symmetries of physical systems, particularly through the application of , which establishes that every of in a corresponds to a . For instance, the U(1) symmetry associated with leads to the conservation of Q, serving as an additive quantum number that remains invariant under phase transformations of the fields. Similarly, translational and rotational symmetries yield conserved linear and angular momenta, respectively, while internal symmetries like those in gauge theories produce additional conserved charges. In the framework of , quantum numbers label the irreducible representations (irreps) of groups that describe these symmetries, providing a mathematical classification of particle states and their transformation properties. For the rotation group SO(3), the orbital angular momentum quantum number l (l = 0, 1, 2, ...) specifies the irreps, with dimension 2l + 1, determining the possible eigenvalues of the . The double cover SU(2) extends this to include half-integer spins, where the j (j = 0, 1/2, 1, ...) labels the irreps, essential for describing both bosonic and fermionic systems in . These representations ensure that states transform consistently under symmetry operations, conserving the corresponding quantum numbers. In , internal symmetries further exemplify this connection. The color symmetry in (QCD) is governed by the non-Abelian gauge group SU(3)_c, where quarks transform in the fundamental triplet representation (dimension 3), carrying , while gluons reside in the adjoint octet representation (dimension 8), mediating the strong interaction. For flavor symmetries, the approximate SU(3) group, proposed in the eightfold way by Gell-Mann and Ne'eman in 1961, classifies light hadrons into multiplets like the octet and decuplet, with quantum numbers such as and labeling the irreps; this symmetry is broken by quark mass differences but remains useful for low-energy phenomenology. In the , the full gauge structure SU(3)_c × SU(2)_L × U(1)_Y assigns particles quantum numbers including color, (with third component T_3 = ±1/2 for left-handed doublets), and Y, where the is given by Q = T_3 + Y/2, ensuring consistency under electroweak transformations. Advanced applications extend this paradigm to conformal groups in (QFT), where the SO(d,2) in d dimensions enlarges the Poincaré , introducing dimensions Δ as quantum numbers that classify irreps and dictate functions in scale-invariant theories like those at critical points. The /CFT correspondence, proposed by Maldacena in , provides a holographic duality between in anti-de Sitter () and a (CFT) on its boundary, linking black hole quantum numbers in the bulk—such as and charges—to CFT s and states, offering insights into and information paradoxes.

Spin-Orbit Coupling and Relativistic Effects

In , the spin-orbit coupling arises as a correction to the non-relativistic , introducing an interaction between the electron's spin angular momentum \mathbf{S} and orbital angular momentum \mathbf{L}. The spin-orbit is given by H_{\mathrm{SO}} = \frac{1}{2m^2 c^2} \frac{1}{r} \frac{dV}{dr} (\mathbf{S} \cdot \mathbf{L}), where m is the , c is the , V(r) is the central potential (e.g., for hydrogen-like atoms), and r is the radial distance. This term originates from the relativistic transformation of the experienced by a moving , leading to splitting in atomic spectra. The j combines the orbital l and s = 1/2, yielding j = l \pm 1/2 (or j_l = |l - s| to l + s). In the for the , the relativistic treatment produces energy levels that depend on j but not on l, resolving the observed in spectral lines. The exact Dirac energies for hydrogen-like atoms scale as E_{n j} = m c^2 \left[1 + \left(\frac{Z \alpha}{n - (j + 1/2) + \sqrt{(j + 1/2)^2 - (Z \alpha)^2}}\right)^2\right]^{-1/2}, where Z is the and \alpha is the , manifesting the j-dependence through the effective . An approximate fine structure formula, derived semi-classically by Sommerfeld in 1919, quantifies the energy shift as \Delta E \approx \frac{\alpha^2}{n^3} \left( \frac{1}{j + 1/2} - \frac{3}{4n} \right) E_n, where E_n is the non-relativistic Bohr energy for n. This expression accurately predicts the splitting for light atoms and laid the groundwork for relativistic . In , the spin-orbit interaction is significantly stronger due to the short-range nuclear forces, playing a crucial role in the proposed by Mayer and Jensen in 1949. It splits nuclear energy levels, such as separating the $1p_{3/2} and $1p_{1/2} subshells in light nuclei like ^{16}\mathrm{O} or ^{17}\mathrm{O}, explaining and stable configurations through enhanced splitting compared to atomic scales. Hyperfine structure emerges from the coupling \mathbf{S} \cdot \mathbf{I} between the electron spin and nuclear spin \mathbf{I}, further splitting levels beyond ; this is modulated by relativistic effects in heavy atoms. The , discovered experimentally in 1947, represents a correction to these levels, shifting the $2S_{1/2} state above the $2P_{1/2} by about 1057 MHz in , arising from interactions. For heavy atoms (Z > 50), the alone inadequately describes quantum numbers due to higher-order effects like and loops, requiring perturbative inclusions for accurate j-dependent energies. Ongoing 2025 precision measurements using muonic atoms, such as those by the MuSEUM collaboration at J-PARC, probe these relativistic modifications at enhanced scales (due to the muon's larger mass), with current hyperfine structure precision of 6.5 ppm in muonic and targets below 0.1 ppm to validate predictions.

References

  1. [1]
    Quantum Numbers and Electronic Structure
    Each solution to a wave equation is characterized by three integers called quantum numbers. Each solution corresponds to a discrete energy and defines a region ...
  2. [2]
    How Quantum Numbers Arise from the Schrodinger Equation
    Quantum numbers arise in the process of solving the Schrodinger equation by constraints or boundary conditions which must be applied to get the solution to fit ...Missing: mechanics | Show results with:mechanics
  3. [3]
    Quantum Numbers and Electron Configurations
    These quantum numbers describe the size, shape, and orientation in space of the orbitals on an atom. The principal quantum number (n) describes the size of the ...Quantum Numbers · Possible Combinations of... · Table: Summary of Allowed...
  4. [4]
    [PDF] Chapter 6 Electronic Structure of Atoms - MSU chemistry
    The principal quantum number, n, describes the energy level on which the orbital resides. • Largest E difference is between E levels. • The values of n are ...<|control11|><|separator|>
  5. [5]
    30.8 Quantum Numbers and Rules – College Physics
    Quantum numbers are used to express the allowed values of quantized entities. The principal quantum number $latex \boldsymbol{n} $ labels the basic states of a ...
  6. [6]
    [PDF] Orbitals & Quantum Numbers
    The first three quantum numbers define the orbital and the fourth quantum number describes the intrinsic electron property called spin. An. 1. 2. − 1. 2 . ms ...
  7. [7]
    [PDF] THE ORIGINS OF THE QUANTUM THEORY
    By the late 1910s physicists had refined Bohr's model, providing a relativistic treatment of the electrons and introducing additional. “quantum numbers.” The ...
  8. [8]
    Atomic and Molecular Quantum Numbers - AstroBaki - CASPER
    Dec 5, 2017 · Each quantum number corresponds to an observable quantity or combination of quantities and describe the quantized states that characterize a system.Missing: definition | Show results with:definition
  9. [9]
    [PDF] Chapter 7. Quantum Theory and Atomic Structure
    An atomic orbital is specified by three quantum numbers. • These determine the orbital's size, shape, and orientation in space. Concept 7-10. The hierarchy of ...
  10. [10]
    [PDF] Notes on Quantum Mechanics - Physics
    Schrödinger equation (equation 1.48 or 1.53), because both involve the same ... quantum numbers nx and ny? What is the lowest level with four-fold ...
  11. [11]
    [PDF] Quantum Mechanics
    • The solutions of the stationary Schrodinger equation are labeled by so called quantum numbers. ... We see that energy takes discrete values: E = h. 2k2. 2m.
  12. [12]
    [PDF] Observables
    From the theoretical point of view, we can consider two commuting observables as a single observable, whose measurement yields two numbers, the value of A and ...
  13. [13]
    [PDF] Additive and Multiplicative Quantum Numbers - UT Physics
    and multiplicative quantum numbers and to investigate the physical implications of stating an empirical conservation law (like “strange- conservation) in a ...
  14. [14]
    [PDF] Isospin
    Discrete symmetries give multiplicative quantum numbers (e.g. parity, charge conjugation). □. Continuous symmetries give additive quantum numbers (e.g. ...
  15. [15]
    [PDF] Chapter 2
    Discrete Symmetry and Multiplicative Quantum Numbers. Space inversion (P), charge-conjugation (C), and time-reversal (T) are examples of discrete symmetry ...
  16. [16]
    [PDF] On common eigenbases of commuting operators
    following definitions: ˆ i) Eigenvalue: A constant λ ∈ С is called an eigenvalue of A if it satisfies the following. equation: ˆ
  17. [17]
    [PDF] CHM 502 – Module 1 – The Postulates of Quantum Mechanics
    2.2 Quantum operators are Hermitian. All valid quantum ... Commuting operators are compatible because there is no uncertainty in a chain of measurements.
  18. [18]
    17 Symmetry and Conservation Laws - Feynman Lectures - Caltech
    The conservation laws are very deeply related to the principle of superposition of amplitudes, and to the symmetry of physical systems under various changes.
  19. [19]
    [PDF] noether.pdf - UCR Math Department
    Nov 9, 2012 · Noether's theorem links the symmetries of a quantum system with its conserved quantities, and is a cornerstone of quantum mechanics. Here we ...
  20. [20]
    [PDF] Elementary Particles and Interactions - MIT
    In addition to such additive quantum numbers, there exist multiplicative quantum num- bers, such as parity that are related to discrete symmetries of ...
  21. [21]
    [PDF] Parity Operator and Eigenvalue
    ☞ Only particles that have all additive quantum numbers = 0 are eigenstates of C. ❍ e.g. γ, ρ, ω, φ, ψ, Υ ❑ These particle are said to be “self conjugate”.
  22. [22]
    Electric Dipole Transitions - Richard Fitzpatrick
    The electric dipole selection rules permit a transition from a $ 2p$ state to a $ 1s$ state of a hydrogen-like atom, but disallow a transition from a $ 2s$ to ...
  23. [23]
    [PDF] 7.1 Dipole Transitions
    These conditions are the selection rules for dipole allowed transitions. Since the dipole interaction Hamiltonian has odd parity, it couples states with ...
  24. [24]
    A.38 Perturbation Theory - Florida State University
    1. Look for good quantum numbers. The quantum numbers that make the energy eigenfunctions of the unperturbed Hamiltonian $H_0$ unique correspond to the ...
  25. [25]
    [PDF] Ay126: LS & JJ coupling, Spectroscopic Terms & Hund's rules
    Apr 11, 2021 · A good approximation can be obtained treat- ing L and S as “good quantum” numbers. This leads to the so-called “L-S” coupling approximation.
  26. [26]
    6.4 Bohr's Model of the Hydrogen Atom - University Physics Volume 3
    Sep 29, 2016 · ... emission lines in this series was discovered in 1885 by Johann Balmer. It is known as the Balmer formula: 1 λ = R H ( 1 2 2 − 1 n 2 ) .
  27. [27]
  28. [28]
    1.3: Bohr's Theory of the Hydrogen Emission Spectrum
    Jul 12, 2019 · The Generalized Rydberg Equation. In an amazing demonstration of mathematical insight, in 1885 Balmer came up with a simple formula for ...
  29. [29]
    Niels Bohr's First 1913 Paper: Still Relevant, Still ... - AIP Publishing
    Nov 1, 2018 · The electron's angular momentum is quantized according to the rule L = nħ. This statement appears first on page 15 of Bohr's paper, where he ...
  30. [30]
    12.2: Bohr Model of the Hydrogen Atom - Physics LibreTexts
    Nov 3, 2024 · One can write the angular momentum quantization condition as. (12.2.11) L = p ⁢ r = n ⁢ h 2 ⁢ π ,. where p is the linear momentum of the ...Missing: principal | Show results with:principal
  31. [31]
    7.21: Zeeman effect - Physics LibreTexts
    Mar 5, 2022 · This is known as the Zeeman effect, discovered in 1896 by the Dutch spectroscopist P. Zeeman. If we start by thinking of an atom with zero ...
  32. [32]
    [PDF] {How Sommerfeld extended Bohr's model of the atom (1913–1916)}
    Jan 30, 2014 · Abstract. Sommerfeld's extension of Bohr's atomic model was moti- vated by the quest for a theory of the Zeeman and Stark effects. The.
  33. [33]
    [PDF] THE ZEEMAN EFFECT - Rutgers Physics
    Zeeman discovered the effect in 1896 and obtained the charge to mass ratio e/m of the ... quantum number and it is usually labeled as S, P, D,F for l=0,1,2,3 ...
  34. [34]
    FIFTY YEARS OF SPIN: Personal reminiscences - Physics Today
    In a one‐page Letter to the Editor of Naturwissenschaften dated 17 October 1925, Samuel A. Goudsmit and I proposed the idea that each electron rotates with an ...
  35. [35]
    [PDF] Stern-Gerlach experiments: past, present, and future
    In 1922 Otto Stern and Walter Gerlach published a paper (Gerlach, 1922) about the experiment that they had just performed at the University of Frankfurt.Missing: m_s | Show results with:m_s
  36. [36]
    [PDF] Rise and premature fall of the old quantum theory - arXiv
    Feb 11, 2008 · Instead of Bohr's ad hoc quantization of angular momentum, Sommerfeld subsequently proposed that atomic process action should be quantized for.Missing: limitations | Show results with:limitations
  37. [37]
    The 1925 Born and Jordan paper “On quantum mechanics”
    Feb 1, 2009 · The ad hoc rules for calculating a quantized energy in the old quantum theory were replaced by a systematic mathematical program. Born and ...Missing: limitations | Show results with:limitations
  38. [38]
    The existence of a neutron - Journals
    Cite this article. Chadwick James. 1932The existence of a neutronProc. R. Soc. Lond. A136692–708http://doi.org/10.1098/rspa.1932.0112. Section. Abstract ...
  39. [39]
    Quantum Numbers for Atoms - Chemistry LibreTexts
    Aug 14, 2024 · A total of four quantum numbers are used to describe completely the movement and trajectories of each electron within an atom.Missing: additive multiplicative
  40. [40]
    [PDF] Multi-electron atoms - Kyle Forinash
    Notation: Capacity of a subshell is 2l(2l+1) due to the exclusion principle. Spectroscopic notation, subshells: s for l = 0; p for l = 1, d for l = 2 etc. f ...
  41. [41]
    2.2.3: Aufbau Principle - Chemistry LibreTexts
    Jan 31, 2024 · The two quantum numbers that are related to energy in multi-electron atoms are \(n\), and \(l\). Thus, orbitals with the lowest values of ...
  42. [42]
    Multi-Electron Atoms - Grandinetti Group
    Aufbau Principle. The energy structure of a many-electron atom is obtained by filling the orbitals one-electron at a time, in order of increasing energy ...
  43. [43]
    11.2: Quantum Numbers of Multielectron Atoms - Chemistry LibreTexts
    May 3, 2023 · The quantum numbers of atoms (and ions) correspond to quantized energy states, called microstates, that depend on the electron configuration.
  44. [44]
  45. [45]
    Hund's Rules
    A set of guidelines, known as Hund's rules, help us determine the quantum numbers for the ground states of atoms.
  46. [46]
    8.8: Term Symbols Gives a Detailed Description of an Electron ...
    Mar 8, 2025 · The Total Orbital Angular Momentum Quantum Number: L. One might naively think that you could get the total angular momentum of an atom ( L ) ...
  47. [47]
    Term Symbols for Atomic Energy Levels - HyperPhysics
    The term uses the multiplicity 2S + 1, total orbital angular momentum L, and total angular momentum J. It assumes that all the spins combine to produce S ...
  48. [48]
    [PDF] Many-Electron Atomic States, Terms, and Levels
    Apr 29, 2008 · (2S + 1 possible values). 2S + 1 = Spin Multiplicity. The total angular momentum is : J = L + S. Jz = Lz + Sz = (ML + MS) ≡ MJ. J = L + S, L + ...
  49. [49]
    9.6: Quantum-Mechanical Orbitals and Electron Configurations
    Feb 24, 2020 · We look at the four quantum numbers for a given electron. Electron configuration notation simplifies the indication of where electrons are ...
  50. [50]
    6.4 Electronic Structure of Atoms (Electron Configurations) - OpenStax
    Feb 14, 2019 · The energy of atomic orbitals increases as the principal quantum number, n, increases. In any atom with two or more electrons, the repulsion ...Missing: multi- | Show results with:multi-
  51. [51]
    [PDF] Chapter Seven
    Quantum Numbers & the Periodic Table. • Principle ... - Specific area of periodic table, spdf “blocks” ... Orbital Filling in Multi-electron Atoms. • Fill ...
  52. [52]
    Electron Configuration - Chemistry LibreTexts
    Jan 29, 2023 · Electron configuration is the standard notation to describe an atom's electronic structure, helping to understand its electron shape and energy.
  53. [53]
    Electronic Configurations Exceptions - Chemistry 301
    Most of the exceptions to the rules occur for the d and f blocks where there is an added stability that comes from having either a half-filled or completely ...<|control11|><|separator|>
  54. [54]
    Predicting nuclear spin - Questions and Answers ​in MRI
    Nuclei with odd numbers of both protons and neutrons have spin quantum numbers that are positive integers. Examples include14N (I=1),2H (deuterium, I=1) ...
  55. [55]
    Nuclear Magnetic Moments and NMR Measurements of Shielding
    Gas phase experiments are applied together with the calculations of shielding in small molecules for the accurate determination of nuclear magnetic moments.
  56. [56]
    [PDF] parity.pdf
    total parity of a state is the product of the intrinsic parities of the particles times the parities of the spatial parts of the wave functions. For example ...
  57. [57]
    Atomic Data for Carbon (C ) - Physical Measurement Laboratory
    Atomic Data for Carbon (C). Atomic Number = 6. Atomic Weight ... 12C, 12.000000, 98.90%, 0. 13C, 13.003355, 1.10%, 1/2, +0.70241. C I Ground State 1s22s22p2 3P0
  58. [58]
    Beta-decay asymmetry and nuclear magnetic moment of
    The asymmetric β-decay radition of 24Nam was used to detect nuclear magnetic resonance signals yielding the magnetic moment |μ( 24 Na m )| = 1.930 (3) n.m. ( ...Missing: angular | Show results with:angular
  59. [59]
    [PDF] 15. Quark Model - Particle Data Group
    May 31, 2024 · Polarized Ω− hyperons have been obtained from polarized Λ and Ξ0 impinging on a nuclear target [102]. In quark models with full isospin symmetry ...
  60. [60]
    [PDF] Tests of Conservation Laws - Particle Data Group
    The lowest-dimension operator containing only SM fields that breaks baryon or lepton number is the d = 5, lepton-number-violating (LNV) 'Weinberg' neutrino ...
  61. [61]
    Symmetries in particle physics: from nuclear isospin to the quark ...
    Apr 24, 2024 · The concept of nuclear isospin is used as a physical motivation to introduce SU(2) and discuss the main techniques of representation theory. ...
  62. [62]
    The role of isospin symmetry in collective nuclear structure - Nature
    In the isospin approach, all nucleons are said to have an isospin t=1/2 (analogous to intrinsic spin s) with neutrons and protons having different projections.
  63. [63]
    Investigation of isospin-symmetry-breaking in mirror energy ... - arXiv
    Mar 20, 2023 · Isospin-symmetry breaking is responsible for the energy difference of excited states in mirror nuclei. It also influences the coefficient of the ...
  64. [64]
    [2303.04917] Strong Interaction Dynamics and Fermi $β$ decay in ...
    Mar 8, 2023 · This paper reviews how strong interactions affect Fermi beta decay, focusing on isospin violation in neutron decay and nuclear short-ranged ...
  65. [65]
    The Connection Between Spin and Statistics | Phys. Rev.
    In the following paper we conclude for the relativistically invariant wave equation for free particles.Missing: URL | Show results with:URL
  66. [66]
    Are all neutrinos left-handed?
    All the neutrinos that we have ever seen are in fact left-handed. Symmetrically, all of the antineutrinos that scientists have ever seen have been right-handed.
  67. [67]
    [1411.4791] Neutrinoless Double-Beta Decay: a Probe of Physics ...
    Nov 18, 2014 · In the Standard Model the total lepton number is conserved. Thus, neutrinoless double-beta decay, in which the total lepton number is violated ...
  68. [68]
    A portrait of the Higgs boson by the CMS experiment ten years after ...
    Jul 4, 2022 · The CMS experiment has observed the Higgs boson in numerous fermionic and bosonic decay channels, established its spin–parity quantum numbers, determined its ...
  69. [69]
    [PDF] Spin determination at the LHC - CERN Indico
    We must determine the spin of any new states to determine the model. Aim to be as model independent as possible. Production distributions are a spin sensitive ...
  70. [70]
    [PDF] Tests of Conservation Laws - Particle Data Group
    Aug 11, 2022 · While strong and electromagnetic forces preserve the quark flavor, the charged-current weak in- teractions generate transitions among the ...
  71. [71]
    The genesis of the CPT theorem | The European Physical Journal H
    May 4, 2022 · In this paper, we want to trace the roots of the CPT theorem, which lie outside the axiomatic approach and culminated with the first explicit ...
  72. [72]
    Phys. Rev. 112, 1375 (1958) - Charge Symmetry of Weak Interactions
    The invariance of strong interactions under G, the product of charge symmetry and charge conjugation, has important consequences for strangeness-conserving ...
  73. [73]
    [PDF] 6. Quantum Electrodynamics - DAMTP
    Noether's theorem told us that these symmetries give rise to conservation laws. Do we now have an infinite number of conservation laws? The answer is no! Gauge ...
  74. [74]
    [PDF] 3 Classical Symmetries and Conservation Laws
    Noether's theorem: For every continuous global symmetry there exists a global conservation law. Figure 3.1 A spacetime 4-volume. Noether theorem reduces to ...
  75. [75]
    [PDF] Quantum Theory, Groups and Representations: An Introduction ...
    This book began as course notes prepared for a class taught at Columbia Uni- versity during the 2012-13 academic year. The intent was to cover the basics of.
  76. [76]
    [PDF] Representation Theory And Quantum Mechanics
    After all, thinking about the previous section, the spin and angular momentum operators are born from representations of SU(2) and. SO(3), respectively.
  77. [77]
    [PDF] Quantum chromodynamics
    Abstract. These lectures provide an overview of Quantum Chromodynamics. (QCD), the SU(3)C gauge theory of the strong interactions. Af-.
  78. [78]
    [PDF] 5 Electroweak Interactions - DAMTP
    We've built the Standard Model around the gauge group G = U(1) ⇥ SU(2) ⇥ SU(3). But it's natural to ask: what are the global symmetries of the Standard Model?Missing: xSU( xU(<|control11|><|separator|>
  79. [79]
    [PDF] 6 Symmetries in Quantum Field Theory - DAMTP
    It's therefore important to understand how symmetry principles arise in QFT, and what their consequences are for the correlation functions and scattering ...
  80. [80]
    [PDF] The AdS/CFT Correspondence arXiv:1501.00007v2 [gr-qc] 19 Feb ...
    Feb 19, 2015 · Counting the black hole microstates has therefore provided an invaluable benchmark for putative theories of quantum gravity. To motivate whence ...
  81. [81]
    [PDF] Origin of the Spin-Orbit Interaction - arXiv
    Abstract. We consider a semi-classical model to describe the origin of the spin-orbit interaction in a simple system such as the hydrogen atom.
  82. [82]
    [PDF] Quantum Physics III Chapter 2: Hydrogen Fine Structure
    Apr 2, 2022 · The dependence on j and absence of dependence on ℓ in the energy shifts could be anticipated from the Dirac equation. The rotation generator ...
  83. [83]
    Mayer-Jensen shell model and magic numbers | Resonance
    Sep 17, 2008 · The addition of a nuclear spin-orbit coupling force to the mean field of the nucleons successfully predicted the nuclear magic numbers and many ...Missing: splitting | Show results with:splitting
  84. [84]
    [2501.02736] Precision measurements of muonium and muonic ...
    Jan 6, 2025 · The MuSEUM collaboration is now performing new precision measurements of the ground state hyperfine structure (HFS) of both muonium and muonic helium atoms.