Fact-checked by Grok 2 weeks ago

Spin quantum number

The spin quantum number, denoted as s, is a fundamental quantum number in quantum mechanics that describes the intrinsic angular momentum, known as spin, possessed by elementary particles such as electrons, quarks, and photons, as well as composite particles like atomic nuclei. This intrinsic spin is a purely quantum mechanical property with no direct classical analog, unlike orbital angular momentum, and it determines both the magnitude of the spin angular momentum operator \mathbf{S}, given by |\mathbf{S}| = \sqrt{s(s+1)} \hbar where \hbar is the reduced Planck's constant, and the possible eigenvalues of its z-component S_z = m_s \hbar, with m_s ranging from -s to +s in integer steps. The value of s can be either an integer (0, 1, 2, ...) for bosons like photons (s = 1) or a half-integer (1/2, 3/2, ...) for fermions like electrons (s = 1/2), fundamentally classifying particles and governing their behavior under the , which prohibits identical fermions from occupying the same . The concept of electron spin was first proposed in 1925 by and to explain the observed in atomic spectra, particularly the anomalous , where they hypothesized that electrons possess an intrinsic of \hbar/2 in addition to their orbital motion. Their seminal paper, "Ersetzung der Hypothese vom unmechanischen Zwang durch eine Forderung bezüglich des inneren Verhaltens jedes einzelnen Elektrons" published in Die Naturwissenschaften, introduced spin as a fourth (m_s = \pm 1/2) alongside the principal (n), azimuthal (l), and magnetic (m_l) quantum numbers, resolving discrepancies in spectral lines that classical models could not account for. This proposal faced initial skepticism from prominent physicists like and due to relativistic concerns about a spinning electron's size, but it was soon validated experimentally through Stern-Gerlach deflection experiments and integrated into the developing framework of . In , the spin quantum number plays a pivotal role in diverse phenomena, from the magnetic properties of materials—where unpaired electron spins contribute to —to , where spin statistics dictate whether particles obey Bose-Einstein or Fermi-Dirac statistics, influencing everything from to the stability of atoms. For instance, the nature of electrons ensures that each can hold at most two electrons with opposite spins, underpinning the building-block structure of the periodic table and enabling technologies like MRI scanners that exploit nuclear spin alignments. Beyond electrons, higher-spin particles like the spin-1 W and Z bosons mediate weak interactions in the , highlighting spin's ubiquity in fundamental forces.

Fundamentals

Definition

The spin quantum number s is a fundamental quantum mechanical property that characterizes the intrinsic angular momentum of elementary particles and composite systems, independent of any spatial motion. Unlike orbital angular momentum, which arises from a particle's position and momentum in space and is quantified by the azimuthal quantum number l, spin is an inherent attribute not associated with classical rotation or orbital dynamics. The magnitude of this intrinsic spin angular momentum is given by \sqrt{s(s+1)} \hbar, where \hbar is the reduced Planck's constant, and s is fixed for each particle type. The possible values of s are either integers (0, 1, 2, ...) or half-integers (1/2, 3/2, ...), leading to a spin multiplicity of $2s + 1, which represents the number of possible orientations of the spin angular momentum along a given axis. For instance, electrons have s = 1/2, resulting in a multiplicity of 2 and thus two possible spin states (often denoted as "up" and "down"), while photons possess s = 1, but only two polarization states corresponding to helicities ±1. Particles with half-integer spin, known as fermions, obey the Pauli exclusion principle, prohibiting identical fermions from occupying the same quantum state, a key factor in the structure of matter.

Nomenclature

In , the intrinsic of a particle is characterized by the s, a non-negative value that determines the magnitude of the spin through the eigenvalue equation \mathbf{S}^2 |s, m_s\rangle = s(s+1) \hbar^2 |s, m_s\rangle. The z-component projection of this spin is described by the quantum number m_s, which takes discrete values m_s = -s, -s+1, \dots, s-1, s. Standard conventions in literature denote the spin angular momentum as the vector \mathbf{S}, with Cartesian components S_x, S_y, and S_z. The eigenvalue of the z-component S_z is given by S_z |s, m_s\rangle = m_s \hbar |s, m_s\rangle, where \hbar is the reduced Planck's constant./10%3A_Pauli_Spin_Matrices/10.1%3A_Spin_Operators The degeneracy of the spin states, or spin multiplicity, is quantified by $2s + 1, representing the number of possible m_s values. This leads to common terminology for specific cases: a state with s = 0 has multiplicity 1 and is termed a singlet; s = 1/2 yields multiplicity 2, called a doublet; and s = 1 gives multiplicity 3, known as a triplet./Electronic_Structure_of_Atoms_and_Molecules/Evaluating_Spin_Multiplicity In , the Russell-Saunders () coupling scheme employs term symbols to label energy levels accounting for spin-orbit interactions. These symbols take the form ^{2S+1}L_J, where uppercase S denotes the total spin quantum number of the electrons, L is the total orbital angular momentum quantum number (represented by letters such as S for L=0, P for L=1, D for L=2, etc.), and J is the arising from the vector sum of L and S./Spectroscopy/Electronic_Spectroscopy/Spin-orbit_Coupling/The_Russell_Saunders_Coupling_Scheme)

Historical Development

Early Concepts

In the late 19th and early 20th centuries, observations of atomic spectra revealed puzzling discrepancies that could not be accounted for by classical orbital alone. The , discovered in 1896, described the splitting of spectral lines in a , but for many atoms, the splitting patterns—known as the anomalous Zeeman effect—deviated from expectations based solely on orbital motion, suggesting an additional source of magnetic interaction. Similarly, measurements of atomic magnetic moments often yielded values that exceeded predictions from orbital contributions, indicating the presence of unexplained intrinsic magnetism in atoms. The Bohr-Sommerfeld model of 1913–1916 extended the Bohr atomic model by incorporating elliptical orbits and relativistic corrections to explain fine structure in hydrogen spectra, yet it failed to fully account for the observed fine splitting in multi-electron atoms or the anomalous Zeeman patterns without invoking ad hoc adjustments. These limitations highlighted the need for an additional angular momentum degree of freedom beyond orbital motion. In 1915, and Wander Johannes de Haas provided early experimental evidence for a coupling between spin-like angular momentum and magnetic moments through their demonstration of rotational effects in demagnetized ferromagnets, where changes in magnetization induced mechanical rotation, consistent with conservation of in atomic-scale currents. By 1919, proposed a theoretical framework for measuring magnetic moments using molecular beams in inhomogeneous fields, serving as a precursor to direct tests of intrinsic quantization. In 1922, Arthur Compton's scattering experiments further implied that s behaved as particles with momentum, raising questions about their internal properties, including potential intrinsic , as classical wave models struggled to explain the results. These pre-1925 anomalies set the stage for the development of , where the concept of electron would resolve many of these inconsistencies.

Key Theoretical Advances

In 1925, and proposed that the possesses an intrinsic , or , with s = 1/2, to account for the anomalous observed in spectra. This idea arose from their analysis of spectral multiplets, where the splitting patterns suggested an additional degree of freedom beyond the orbital , initially modeled as a classical spinning top with a of one . Although Uhlenbeck and Goudsmit initially hesitated due to concerns about unphysically high rotation speeds and radiation losses implied by a classical interpretation, they quantized the motion, assigning it the \hbar/2 based on the . Their was influenced by Pauli's earlier 1925 exclusion principle, which required a fourth for to explain , prompting the interpretation of this degree as . A key resolution to classical inconsistencies in the Uhlenbeck-Goudsmit model came in 1926 from , who introduced the concept of —a relativistic kinematic effect arising from the 's motion in an . This halves the expected spin-orbit coupling energy, yielding a gyromagnetic ratio, or g-factor, of 2 for the , consistent with the observed in spectra without invoking adjustments. 's calculation demonstrated that the relativistic between the lab frame and the accounts for the factor of 1/2 in the spin-orbit interaction, aligning the theoretical with experimental values. In 1927, Pauli provided a formal quantum mechanical framework for spin-1/2 particles by extending Schrödinger's wave mechanics to include the magnetic . He introduced a two-component and the three \sigma_x, \sigma_y, \sigma_z, which serve as operators for the spin components, satisfying the commutation relations [\sigma_i, \sigma_j] = 2i \epsilon_{ijk} \sigma_k. These matrices enabled the description of spin-orbit interactions and the within non-relativistic , with the incorporating the electron's . The anticommutation relations \{ \sigma_i, \sigma_j \} = 2 \delta_{ij} I emerged naturally from this representation, laying the groundwork for fermionic statistics. The culmination of these advances occurred in 1928 when Paul Dirac developed a relativistic wave equation for the electron, inherently incorporating spin as a consequence of Lorentz invariance and linearity in the Dirac matrices. Dirac's formulation resolved remaining inconsistencies in the non-relativistic treatments, such as the fine structure and the g-factor of 2, by predicting both positive and negative energy solutions for spin-1/2 particles. This theory unified quantum mechanics with special relativity, establishing spin as a fundamental intrinsic property rather than an add-on.

Spin in Elementary Particles

Electron Spin

The electron possesses an intrinsic angular momentum characterized by the spin quantum number s = \frac{1}{2}, a fundamental property proposed by and in 1925 to explain the of atomic spectra. This half-integer value distinguishes electrons as fermions and gives rise to two possible projections of the spin angular momentum along a quantization axis, denoted by the magnetic spin quantum number m_s = \pm \frac{1}{2}, conventionally referred to as spin-up (m_s = +\frac{1}{2}) and spin-down (m_s = -\frac{1}{2}) states. These states represent the only allowed orientations for an electron's spin in a given , reflecting the quantized nature of spin as an internal degree of freedom rather than classical . The spin of the electron generates a magnetic dipole moment, given by \boldsymbol{\mu} = -g \mu_B \frac{\mathbf{S}}{\hbar}, where \mu_B is the Bohr magneton, \mathbf{S} is the spin angular momentum operator, and g \approx 2 is the electron's spin g-factor, which arises naturally from the Dirac relativistic quantum theory of the electron formulated in 1928. This near-exact value of 2 (with small quantum electrodynamic corrections) implies that the magnetic moment is twice as large as expected from a simple orbital analogy, leading to significant interactions in external fields. Additionally, the spin couples with the electron's orbital motion through the spin-orbit interaction, producing an effective magnetic field in the electron's rest frame that shifts energy levels; the interaction Hamiltonian is proportional to \mathbf{L} \cdot \mathbf{S}, where \mathbf{L} is the orbital angular momentum, and this relativistic effect fine-tunes atomic spectra and influences selection rules in transitions. The , formulated by in 1925, states that no two in an atom can occupy the same , meaning they cannot share identical values for the principal n, orbital quantum number l, m_l, and spin quantum number m_s. This principle, directly tied to the s = \frac{1}{2} nature of spin, enforces that each spatial orbital can hold at most two electrons with opposite spins, which underpins the building of electron shells and subshells, thereby explaining the structure of the periodic table and the chemical properties of elements./University_Physics_III_-Optics_and_Modern_Physics(OpenStax)/08%3A_Atomic_Structure/8.05%3A_The_Exclusion_Principle_and_the_Periodic_Table) In chemical contexts, electron spin plays a key role in molecular and reactivity; atoms or molecules with unpaired electrons exhibit due to the net from aligned spins, as seen in complexes and oxygen./Physical_Properties_of_Matter/Atomic_and_Molecular_Properties/Magnetic_Properties) Free radicals, such as the (OH•), possess one or more unpaired electrons, making them highly reactive species that drive processes like and , with their spin states influencing bond formation and stability. Experimentally, the electron's spin response to a magnetic field is quantified by its electron spin resonance (ESR) frequency, approximately 28 GHz in a 1 T field, corresponding to the \gamma_e / 2\pi \approx 28.025 GHz/T for a free electron.

Nuclear and Other Particle Spins

The nuclear spin quantum number I characterizes the intrinsic of atomic nuclei, arising from the spins of constituent protons and neutrons, each with spin i = 1/2. For composite nuclei, I depends on pairing: even-even nuclei (even protons and even neutrons) typically have I = 0, as in ^{12}\mathrm{C}, while odd numbers yield or integer values, such as I = 1/2 for ^{13}\mathrm{C} due to its seven neutrons./Spectroscopy/Magnetic_Resonance_Spectroscopies/Nuclear_Magnetic_Resonance/NMR_-_Theory)/19%3A_Nuclear_Magnetic_Resonance_Spectroscopy/19.01%3A_Theory_of_Nuclear_Magnetic_Resonance) Nuclear spin I > 0 produces a moment that interacts with the from electrons, causing hyperfine splitting in spectral lines and fine details in levels. This effect is observable in spectra and enables techniques like (NMR) for nuclei with I > 0, which aligns spins in a to reveal molecular environments through resonance frequencies. For example, (^2\mathrm{H}) with I = 1 exhibits an electric moment from its prolate charge distribution, influencing NMR line shapes via quadrupolar interactions./19%3A_Nuclear_Magnetic_Resonance_Spectroscopy/19.01%3A_Theory_of_Nuclear_Magnetic_Resonance)/Spectroscopy/Magnetic_Resonance_Spectroscopies/Nuclear_Magnetic_Resonance/NMR_-_Theory/NMR_Interactions/Quadrupolar_Coupling) In , spins extend to fundamental constituents beyond nucleons. Protons and neutrons, as baryons, have total spin s = 1/2, composed of three quarks each with s = 1/2. Gluons, the force carriers of the strong interaction, possess spin s = 1, while the is a scalar particle with spin s = 0. These assignments align with the classification. The spin value determines particle statistics via the spin-statistics theorem: half-integer spins (s = 1/2, 3/2, \ldots) classify particles as fermions, obeying Fermi-Dirac statistics and the , as seen in quarks and nucleons. Integer spins (s = 0, 1, 2, \ldots) denote bosons, following Bose-Einstein statistics and permitting Bose-Einstein condensation, exemplified by gluons and the . This connection underpins and matter's quantum behavior./19%3A_Atoms/19.01%3A_Fermions_and_Bosons)

Experimental Observation

Stern-Gerlach Experiment

The Stern-Gerlach experiment, conducted in 1922 by and at the University of , provided the first direct experimental evidence for the quantization of atomic magnetic moments, a key prediction of early . The setup involved evaporating silver atoms from an oven heated to approximately 1000°C, forming a through narrow slits (about 0.03 mm wide) to ensure a well-defined path. This beam then passed through an inhomogeneous generated by an with pole pieces producing a of around 0.1 and a gradient of about 10 tesla per centimeter over a 3.5 cm length. Silver was selected because its ground-state configuration features a single unpaired , resulting in a net dominated by this electron's properties, simplifying the interpretation. Classically, was expected to produce a continuous spread of deflections on a detector screen, as the random orientations of magnetic moments relative to would cause a range of forces and thus a smeared line rather than spots. However, the observed results defied this prediction: the silver beam split into two distinct spots separated by approximately 0.2 mm on the detector plate, corresponding to upward and downward deflections. This indicated that the atoms' magnetic moments could only in two directions along , with deflections proportional to the m_j = \pm 1/2. The interpretation of these results established the quantization of the spin , refuting classical models of continuously varying vectors and confirming Niels Bohr's concept of space quantization. The measured was approximately one Bohr magneton, aligning with expectations for a where the m_s = \pm 1/2 determines the force F_z = -\mu_z \frac{\partial B_z}{\partial z}, with \mu_z being the z-component of the . This experiment later became foundational for understanding electron spin, as the in silver atoms primarily governs the observed effect. The experiment's success relied on precise apparatus alignment, with slit tolerances under 0.01 mm, and was limited to paramagnetic atoms like silver that possess a net magnetic moment; diamagnetic atoms, with paired electrons and no such moment, would show no deflection.

Spectroscopic Methods

Spectroscopic methods measure spin quantum numbers indirectly by detecting energy transitions between spin states induced by external magnetic fields and electromagnetic radiation. These techniques exploit the Zeeman effect, where spin angular momentum interacts with the field to split energy levels, allowing resonance absorption when the radiation frequency matches the splitting energy. The general resonance condition is h\nu = g \mu B, where h is Planck's constant, \nu the radiation frequency, g the gyromagnetic ratio, \mu the magnetic moment, and B the magnetic field strength; this makes the methods highly sensitive to local magnetic environments, including those from nearby atoms or defects. Electron Paramagnetic Resonance (EPR), also called Electron Spin Resonance (ESR), detects unpaired electrons in paramagnetic materials through absorption of microwave radiation in a static . The energy splitting between spin states follows \Delta E = g \mu_B B, with g the electron g-factor (approximately 2 for free electrons), \mu_B the , and B the field; resonance occurs when microwaves bridge this gap, yielding spectra that reveal spin concentrations, interactions, and dynamics. EPR was discovered in 1945 by Evgenii Zavoisky, who observed paramagnetic absorption in salts. Today, it is routinely applied in to characterize defects, ions, and free radicals in solids, providing insights into electronic structure and reactivity. Nuclear Magnetic Resonance (NMR) spectroscopy probes nuclear spins, such as those of hydrogen or , using radiofrequency pulses in a to induce transitions between aligned and anti-aligned states. Spectra display chemical shifts, which quantify the influence of local shielding the nuclear spin from the external field, and J-couplings, which arise from indirect through-bond interactions between neighboring nuclear spins mediated by electrons. NMR was independently developed by and Edward Purcell in 1946, earning them the for revealing nuclear magnetic moments. In , it elucidates molecular structures, dynamics, and bonding in complex systems like polymers and catalysts. Other techniques include , which observes nuclear transitions through recoilless emission and absorption of gamma rays from excited nuclei, offering high-resolution probes of hyperfine interactions and states in solids, especially for iron-57. (ODMR) enhances sensitivity by optically exciting systems and detecting via changes in , commonly used for shallow spins in semiconductors and defects. These methods collectively enable precise, non-destructive characterization of properties across atomic, molecular, and solid-state scales.

Mathematical Framework

Relation to Spin Angular Momentum

The spin quantum number s labels the intrinsic associated with elementary particles and composite systems, formalized through the spin angular momentum operator \mathbf{S}. This operator acts on the of the particle's spin degrees of freedom, with eigenvalues determined by s. The magnitude of the spin angular momentum vector is \sqrt{s(s+1)} \hbar, reflecting the quantum mechanical quantization of its squared magnitude \mathbf{S}^2 = s(s+1) \hbar^2. The component along any chosen quantization axis, conventionally the z-axis, is S_z = m_s \hbar, where m_s takes values from -s to +s in steps. The spin operators S_x, S_y, and S_z obey the fundamental commutation relations of angular momentum algebra: [S_x, S_y] = i \hbar S_z, with cyclic permutations for the other pairs. These relations ensure that \mathbf{S} behaves as a vector operator under rotations, paralleling the structure of orbital angular momentum \mathbf{L} but arising intrinsically rather than from spatial degrees of freedom. For spin-1/2 particles like the electron, s = 1/2, yielding m_s = \pm 1/2 and a magnitude \sqrt{3/4} \hbar \approx 0.866 \hbar, distinct from the z-component projections of \pm \hbar/2. In the semiclassical interpretation, the expectation value \langle \mathbf{S} \rangle for a state with definite m_s precesses on a of \sqrt{s(s+1)} \hbar, but for large s, such as in spins, the discreteness fades, and \langle |\mathbf{S}| \rangle \approx s \hbar, approaching a classical of fixed length s \hbar with continuous orientation. This quantum-to-classical transition highlights the underlying vector nature while preserving the non-commutativity that forbids simultaneous precise knowledge of all components. However, the inherent discreteness in m_s ensures that measurements yield quantized projections, underscoring the quantum origin even in high-spin limits. When subjected to an external \mathbf{B} along the z-axis, the Hamiltonian H = - \boldsymbol{\mu} \cdot \mathbf{B}, with \boldsymbol{\mu} = - [g](/page/G) \mu_B \mathbf{S} / \hbar (where \mu_B is the and g \approx 2 for electrons), induces equivalent to of \langle \mathbf{S} \rangle around \mathbf{B} at the Larmor frequency \omega = [g](/page/G) \mu_B B / \hbar. For electrons, this yields \omega \approx 1.76 \times 10^{11} rad/s per , enabling phenomena like . This arises from the accumulation in the components, maintaining the fixed magnitude while rotating the value .

Algebraic Representation

The algebraic structure of the spin quantum number is captured by the Lie algebra of the SU(2), whose generators are the spin operators S_x, S_y, and S_z. These operators satisfy the commutation relations [S_x, S_y] = i \hbar S_z and cyclic permutations thereof, which define the (2) . The operator \mathbf{S}^2 = S_x^2 + S_y^2 + S_z^2 commutes with each generator, [\mathbf{S}^2, S_i] = 0, and in any labeled by the spin quantum number s, it acts as the scalar multiple \mathbf{S}^2 = s(s+1) \hbar^2 I, where I is the identity on the (2s+1)-dimensional . These finite-dimensional s of (2) are unique up to unitary equivalence for each or s \geq 0, with the $2s+1 determining the number of possible magnetic quantum numbers m = -s, -s+1, \dots, s. For the fundamental case of spin s = 1/2, the representation is two-dimensional, and the spin operators take the explicit form S_i = \frac{\hbar}{2} \sigma_i (for i = x, y, z), where the \sigma_i are the introduced by Pauli to describe the magnetic properties of the . The are: \sigma_x = \begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix}, \quad \sigma_y = \begin{pmatrix} 0 & -i \\ i & 0 \end{pmatrix}, \quad \sigma_z = \begin{pmatrix} 1 & 0 \\ 0 & -1 \end{pmatrix}. Each \sigma_i is Hermitian (\sigma_i^\dagger = \sigma_i), unitary (\sigma_i^{-1} = \sigma_i), and satisfies \sigma_i^2 = I, with the property that the product of any two distinct matrices is the third up to a factor of i (e.g., \sigma_x \sigma_y = i \sigma_z). These relations ensure the su(2) algebra is realized faithfully in this representation. To navigate the basis states |s, m\rangle (eigenstates of \mathbf{S}^2 and S_z), one introduces the raising and lowering operators S_\pm = S_x \pm i S_y. These satisfy [S_z, S_\pm] = \pm \hbar S_\pm and act as S_+ |s, m\rangle = \hbar \sqrt{s(s+1) - m(m+1)} \, |s, m+1\rangle and S_- |s, m\rangle = \hbar \sqrt{s(s+1) - m(m-1)} \, |s, m-1\rangle, with vanishing action at the boundaries m = \pm s. For s=1/2, this yields the explicit transitions between the states |+\rangle and |-\rangle, such as S_+ |-\rangle = \hbar |+\rangle.

Relativistic and Composite Systems

Spin in the Dirac Equation

The , formulated by in 1928, provides a relativistic wave equation for the that naturally incorporates spin as an intrinsic property, resolving inconsistencies in earlier non-relativistic with . This equation addressed the "duplexity" observed in atomic spectra, where the number of stationary states was twice that predicted without spin, by deriving a linear in both and time that is . The requirement for Lorentz invariance in a quantum mechanical description of particles leads to the use of 4-component wave functions, as the spinor representation of the accommodates the spin degrees of freedom. The Dirac equation in its standard form is given by i \hbar \frac{\partial \psi}{\partial t} = c \boldsymbol{\alpha} \cdot \mathbf{p} \psi + \beta m c^2 \psi, where \psi is a 4-component spinor, \mathbf{p} = -i \hbar \nabla is the momentum operator, m is the electron mass, c is the speed of light, \hbar is the reduced Planck's constant, and \boldsymbol{\alpha} and \beta are 4×4 matrices satisfying specific anticommutation relations to ensure the equation's relativistic invariance. These matrices, along with the Pauli matrices embedded in the spinor structure, encode the spin angular momentum, making spin an emergent feature rather than an ad hoc addition. Solutions to the free-particle Dirac equation yield positive-energy states with energies E = +\sqrt{(pc)^2 + (mc^2)^2} and negative-energy states with E = -\sqrt{(pc)^2 + (mc^2)^2}, each doubly degenerate due to the two possible spin projections (up and down) along a quantization axis. In the Dirac sea interpretation, the positive-energy solutions describe electrons with spin up and down, while the negative-energy solutions, upon filling the sea and considering holes, correspond to positrons with opposite spin. A key prediction of the Dirac equation is the electron's , or g-factor, of exactly 2, arising naturally from the relativistic coupling of to the orbital motion without additional assumptions. This value emerges in the term with electromagnetic fields, where the magnetic moment is \mu = -g \frac{e}{2m} \mathbf{S} with g=2 and \mathbf{S} the , matching experimental observations for the electron's and explaining the splitting in spectra. To connect the to non-relativistic , the Foldy-Wouthuysen transformation is applied, which performs a unitary on the to decouple the positive- and negative-energy components and separate the large (upper) and small (lower) components of the in the low-velocity limit. This transformation reveals the for the positive-energy sector, where the operators appear explicitly as the acting on the 2-component , confirming as \mathbf{S} = \frac{\hbar}{2} \boldsymbol{\sigma}, and yields relativistic corrections including the g=2 term. The transformation preserves the Lorentz invariance of the original while facilitating the interpretation of in everyday non-relativistic contexts.

Total Spin in Atoms and Molecules

In multi-electron atoms, the total \mathbf{S} arises from the vector sum of individual spins, \mathbf{S} = \sum_i \mathbf{s}_i, where each \mathbf{s}_i has magnitude \hbar \sqrt{s(s+1)} with s = 1/2. This total couples with the total \mathbf{L} = \sum_i \mathbf{l}_i to form the total \mathbf{J} = \mathbf{L} + \mathbf{S}, a scheme known as Russell-Saunders or LS coupling, which is predominant in light atoms where spin-orbit interactions are weak compared to electrostatic correlations. In this approximation, the quantum numbers S (total ) and L (total orbital) are good labels for atomic states, with J determined by addition rules. For example, in the atom's excited states, the ground configuration $1s^12s^1 yields a (S=0) with symmetric spatial wavefunction and a (S=1) with antisymmetric spatial wavefunction, the latter lower in energy due to reduced electron-electron repulsion from greater average separation. For heavier atoms, where spin-orbit coupling dominates, the jj coupling scheme applies, in which each electron's spin and orbital momenta first couple to form individual j_i = l_i + s_i, and these j_i then sum to total J = \sum_i j_i. This regime is relevant for elements beyond (Z > 40), as relativistic effects enhance spin-orbit splitting, invalidating pure LS coupling. In transition metals, d-electron configurations often exhibit high-spin states following , which maximize S by aligning unpaired spins parallel in degenerate orbitals to minimize exchange energy; for instance, in octahedral \mathrm{d}^5 complexes like \mathrm{Mn}^{2+}, the high-spin S=5/2 state has five unpaired electrons. In molecules, total spin determination follows analogous principles, with predicting ground-state multiplicities by maximizing S and L for equivalent electrons. The oxygen molecule \mathrm{O_2}, with (\pi^*_{2p})^2, has a (S=1) due to two unpaired \pi^* electrons with parallel spins, as dictated by the first Hund's rule, leading to observable in susceptibility measurements. further refines energy levels through coupling of the total electronic \mathbf{J} with nuclear spin \mathbf{I}, forming \mathbf{F} = \mathbf{I} + \mathbf{J}; this interaction, first quantified by Fermi, splits spectral lines by amounts proportional to the product of electronic and nuclear g-factors.

References

  1. [1]
    Properties of Spin Angular Momentum - Richard Fitzpatrick
    Spin angular momentum has similar properties to orbital, but its quantum number can be half-integer, unlike orbital, and is purely quantum mechanical.
  2. [2]
    30.8 Quantum Numbers and Rules – College Physics
    Quantum numbers express quantized values of entities like energy and angular momentum. Key numbers include principal, angular momentum, spin, and spin ...
  3. [3]
    [PDF] A Short History of Spin - arXiv
    Nov 20, 2013 · and Goudsmit first hypothesized [6] intrinsic spin as a property of the electron. This occurred over the strong objections of some prominent ...<|control11|><|separator|>
  4. [4]
  5. [5]
    Quantum Numbers and Electron Configurations
    This is called the spin quantum number (s) because electrons behave as if they were spinning in either a clockwise or counterclockwise fashion. One of the ...
  6. [6]
    Electron spin - HyperPhysics Concepts
    An electron spin s = 1/2 is an intrinsic property of electrons. Electrons have intrinsic angular momentum characterized by quantum number 1/2.
  7. [7]
    [PDF] Chapter 11. Angular Momentum: General Theory
    In addition to the orbital angular momentum, the particles studied in quantum mechanics have an intrinsic angular momentum called spin. This is not caused ...
  8. [8]
    The Facts - UCR Math Department
    The same result holds for the component of angular momentum measured along any axis. The total magnitude of angular momentum is given by $\sqrt{j_x^2+j_y^2+j_z^ ...
  9. [9]
    [PDF] chapter 11 spin
    Real particles have “intrinsic” angular momentum. This degree of freedom is called spin. It is a fixed property of a particle that does not change.
  10. [10]
    Pauli Exclusion Principle - HyperPhysics
    The Pauli exclusion principle is part of one of our most basic observations of nature: particles of half-integer spin must have antisymmetric wavefunctions, and ...
  11. [11]
    4.3 LS and jj Coupling Schemes - Atomic Physics - Fiveable
    Term symbol for an electronic state in LS coupling is written as $^{2S+1}L_J$. $S$ is total spin angular momentum quantum number; $L$ is total orbital angular ...
  12. [12]
    One hundred years ago Alfred Landé unriddled the Anomalous ...
    In order to commemorate Alfred Landé's unriddling of the anomalous Zeeman Effect a century ago, we reconstruct his seminal contribution to atomic physics.Missing: Erklärung | Show results with:Erklärung
  13. [13]
    [PDF] Rise and premature fall of the old quantum theory - arXiv
    Feb 11, 2008 · fails to explain the fine structure. Therefore Sommerfeld refined ... history of the hydrogen molecule ion and Pauli's failure to solve it with ...Missing: sources | Show results with:sources
  14. [14]
    [PDF] On the history of the Einstein-de Haas effect - Physics - UMD
    However, the spin component of the magnetic moment plays the domi- nant role in ferromagnetic materials on which the experi- ments of Refs. 9 and 16 and ...
  15. [15]
    THE STERN-GERLACH EXPERIMENT - Europhysics News
    The historic Stern-Gerlach experiment (SGE), which was performed in 1922 in Frankfurt, is reviewed from an experimental point of view.Missing: precursor | Show results with:precursor
  16. [16]
    Arthur Compton and the mysteries of light - Physics Today
    Dec 1, 2022 · Compton's scattering results resolved long-standing controversies regarding the nature of free radiation and rescued Albert Einstein's long- ...
  17. [17]
    The Vector Model and the Pauli Principle | Phys. Rev.
    Abstract. The vector model is examined in the light of the quantum theory of wave fields. First the distributions of electrons among the electronic states are ...Missing: original paper
  18. [18]
    The quantum theory of the electron - Journals
    To meet the difficulty, Goudsmit and Uhlenbeck have introduced the idea of an electron with a spin angular momentum of half a quantum and a magnetic moment of ...
  19. [19]
    Spinning Electrons and the Structure of Spectra - Nature
    The idea of a quantised spinning of the electron was put forward for the first time by AK Compton (Journ. Frankl. Inst., Aug. 1921, p. 145).
  20. [20]
    Quantum Numbers in Multielectron Atoms
    ... quantum number for electron spin, \( m_s \). There are two values which we call "spin up" or "spin down". These have values of +1/2 and -1/2. In general for ...
  21. [21]
    [PDF] Spins in Magnetic Fields
    Approximating the g-factor by 2, the electron magnetic moment can be written,. µ = −geµB σ. 2. ≈ −µBσ. (19). The minus sign in this equation means that the ...
  22. [22]
    Spin-Orbit Interaction - an overview | ScienceDirect Topics
    Spin-orbit interaction (SOI) refers to the coupling between an electron's spin and its orbital motion around a nucleus, resulting from the relativistic ...
  23. [23]
    [PDF] On the Connexion between the Completion of Electron Groups in an ...
    Apr 16, 2010 · PAULI. Z. Physik 31, 765ff (1925). Especially in connexion with Millikan and Landé's observation that the alkali doublet can be represented ...
  24. [24]
    EPR Spectroscopy — Free Radicals Chemistry Group
    When the spin of the odd electron of a free radical is placed in a magnetic field, it may have two orientations, one parallel and the other antiparellel to the ...
  25. [25]
    electron gyromagnetic ratio in MHz/T
    electron gyromagnetic ratio in MHz/T $\gamma_{\rm e}$. Numerical value, 28 024.951 3861 MHz T-1. Standard uncertainty, 0.000 0087 MHz T-1.
  26. [26]
    NMR Spectroscopy - MSU chemistry
    Examples are I = 1 ( 2H, 14N ). Even mass nuclei composed of even numbers of protons and neutrons have zero spin ( I = 0 ). Examples are 12C, and 16O.Missing: quantum | Show results with:quantum
  27. [27]
    Hyperfine Structure
    The interaction between the spin of the nucleus and the angular momentum of the electron causes a further (hyperfine) splitting of atomic states.
  28. [28]
    The Standard Model - Summary - The Physics Hypertextbook
    Summary · are the particles that mediate interactions · obey Bose-Einstein statistics · have integral spin quantum numbers (±0, ±1, ±2, …).
  29. [29]
    The Standard Model of Particle Physics and Beyond
    The Higgs boson (H), a scalar particle with spin-0, arises from the Higgs field, imparting mass to other particles through the Higgs mechanism. Fig. 1: ...
  30. [30]
    [PDF] Spin–Statistics Theorem - UT Physics
    The spin–statistics theorem says that the fields of integral spins commute (and therefore must be quantized as bosons) while the fields of half-integral spin ...
  31. [31]
  32. [32]
    Stern and Gerlach: How a Bad Cigar Helped Reorient Atomic Physics
    Dec 1, 2003 · The history of the Stern–Gerlach experiment reveals how persistence, accident, and luck can sometimes combine in just the right ways.
  33. [33]
    100 Years Ago, a Quantum Experiment Explained Why We Don't ...
    Feb 8, 2022 · In Stern and Gerlach's experiment, the physicists detected the splitting of the beam, which they saw as confirmation of Bohr's odd prediction: ...
  34. [34]
    Zeeman – Electron Paramagnetic Resonance | ETH Zurich
    The splitting between the two energy states is called electron Zeeman interaction (EZI) and is proportional to the magnitude of B0, as illustrated in Figure 1.
  35. [35]
    Zavoisky and the discovery of EPR | Resonance
    Dec 17, 2015 · E K Zavoisky. Paramagnetic Relaxation of Liquid Solutions for Perpendicular Fields, J. Phys., Vol.9, pp.211–216, 1945. Received July 12, ...Missing: seminal paper
  36. [36]
    On the Role and Applications of Electron Magnetic Resonance ...
    Mar 16, 2021 · Electron Paramagnetic Resonance, EPR, has been employed in experimental research on surface chemistry and heterogeneous catalysis since the beginning of the ...
  37. [37]
    An Introduction to Biological NMR Spectroscopy - PMC
    Scalar coupling (or J-coupling) is a through-bond interaction between two nuclei (A and X) with a nonzero spin. It is an indirect interaction between the two ...
  38. [38]
    Discovery of Nuclear Magnetic Resonance: Rabi, Purcell, and Bloch
    Apr 10, 2020 · After a brief review of background information on the quantum concept of spin and its relative, magnetic moment, we examine the detection of NMR ...Missing: seminal | Show results with:seminal
  39. [39]
    Mössbauer Spectroscopy - SERC (Carleton)
    Jun 15, 2018 · The technique of Mössbauer spectroscopy is widely used in mineralogy to examine the valence state of iron.
  40. [40]
    Optical detection of magnetic resonance - PMC - PubMed Central
    Magnetic resonance spectroscopy basically measures the interaction of electronic or nuclear angular momenta with each other and with external magnetic fields.
  41. [41]
    [PDF] Ersetzung der Hypothese vom unmechanischen Zwang durch eine ...
    G. E. UHLENBECK und S. GOUDSMIT. Leiden, den 17. Oktober 1925. Instituut voor Theoretische Natuurkunde. Es ist nfir ein Bedfirfnis, festzustellen, dab Prof ...Missing: Erklärung | Show results with:Erklärung
  42. [42]
    [PDF] On the quantum mechanics of magnetic electrons
    Phys. 43 (1927), 601-623. On the quantum mechanics of magnetic electrons. By W. PAULI, Jr ...
  43. [43]
    [PDF] Introduction to Quantum Spin Systems - Lecture 4: SU(2)
    Jan 24, 2011 · The Lie algebra su2 is defined as the tangent space to SU(2) at the identity. Page 4. 7. The previous calculation shows that the spin-1/2 ...
  44. [44]
    [PDF] A tutorial for SU(2) and spin waves
    Jun 25, 2014 · The Lie group SU(2) is the symmetry group of the quantum spin. For example, the electron carries a spin with magnitude 1/2 in the physics ...
  45. [45]
    [PDF] Theory of Angular Momentum and Spin
    The theory involves rotational symmetry transformations, the group SO(3), and spin-1 states (SU(2)). Spin states behave similarly to angular momentum states.
  46. [46]
    [PDF] The Dirac Equation and the Lorentz Group - Physics Courses
    Feb 16, 2009 · This then shows that the Dirac equation represents a particle with conserved total angular momentum. J = L + S and with spin equal to 1/2. Now ...
  47. [47]
    [PDF] 4. The Dirac Equation - DAMTP
    However, for spinor fields, the magic of the µ matrices means that the Dirac Lagrangian is Lorentz invariant. The Dirac equation mixes up different components ...
  48. [48]
    Dirac Plane Wave Solution
    Dirac Plane Wave Solution. We now have simple solutions for spin up and spin down for both positive energy and ``negative energy'' particles at rest.
  49. [49]
    [PDF] Introduction to the Dirac Equation
    ... Dirac, who in 1928 published his work on his wave equation, which we ... as far as anyone knew the Dirac equation gave the exact g-factor of the electron.
  50. [50]
    None
    Nothing is retrieved...<|separator|>
  51. [51]
    [PDF] The Foldy-Wouthuysen Transformation
    This transformation block-diagonalizes the matrices that appear in the Dirac equation, thereby decoupling the upper and lower 2-component spinors of the 4- ...
  52. [52]
    Atomic Spectroscopy - Different Coupling Schemes | NIST
    Oct 3, 2016 · LS Coupling (Russell-Saunders Coupling). Some of the examples given below indicate notations bearing on the order of coupling of the electrons.
  53. [53]
    Helium Atom - Richard Fitzpatrick
    Incidentally, helium in the spin singlet state is known as para-helium, whereas helium in the triplet state is called ortho-helium. As we have seen, for the ...
  54. [54]
    Russell-Saunders or LS Coupling - HyperPhysics
    In L-S coupling, orbital and spin angular momenta combine to form total angular momentum (J=L+S). This is common in light atoms, while j-j coupling is for  ...
  55. [55]
    Spinning around in Transition-Metal Chemistry - ACS Publications
    Nov 14, 2016 · In transition-metal complexes, the electronic spin can manifest itself as unpaired electrons at the metal; however, there also exist redox ...
  56. [56]
    [PDF] Energetic and chemical reactivity of atomic and molecular oxygen.
    Jun 28, 2010 · 1 The multiplicity is given by 2S+1, where S is the spin. The spin of an electron is (+/-) 1/2. 3. P. 1. S. 1. D. Energetic and chemical ...
  57. [57]
    Theory of Hyperfine Structure Separations | Phys. Rev.
    Expressions are derived for the hyperfine structure separations of the levels of complicated electron configurations in different types of coupling.