Fact-checked by Grok 2 weeks ago

Total angular momentum quantum number

The total angular momentum quantum number, denoted as j (for a single particle) or J (for a composite system such as an ), quantifies the magnitude of the total vector \mathbf{J} = \mathbf{L} + \mathbf{S}, where \mathbf{L} is the orbital and \mathbf{S} is the spin . In quantum mechanics, this quantum number arises from the addition of angular momenta, where the possible values of j range from |l - s| to l + s in integer steps of 1, with l being the orbital angular momentum quantum number and s the spin quantum number (typically s = 1/2 for electrons). For example, in a hydrogen-like atom, an electron in an orbital with l = 1 yields j = 1/2 or j = 3/2. For multi-electron atoms, J is determined through coupling schemes like Russell-Saunders (L-S) coupling, where the total orbital angular momentum L (sum of individual l_i) and total spin S (sum of individual s_i) are first coupled separately before combining to form J, with allowed J values again spanning |L - S| to L + S. This coupling is crucial for lighter elements, while heavier atoms may require jj-coupling due to stronger spin-orbit interactions. The eigenvalues associated with J include the magnitude squared J^2 = \hbar^2 J(J+1) and the z-component J_z = m_J \hbar, where m_J takes $2J + 1 discrete values from -J to +J in steps of 1, leading to a degeneracy of $2J + 1 for each J state. These properties underpin the in atomic spectra and the classification of atomic terms using symbols like ^{2S+1}L_J.

Angular Momentum in Quantum Mechanics

Orbital Angular Momentum

In quantum mechanics, the orbital angular momentum of a particle arises from its motion in a central potential and is represented by the operator \mathbf{L} = \mathbf{r} \times \mathbf{p}, where \mathbf{r} is the position operator and \mathbf{p} = -i\hbar \nabla is the linear momentum operator. This vector operator quantizes the rotational degrees of freedom, leading to discrete eigenvalues rather than continuous classical values. For a particle in a spherically symmetric potential, such as the Coulomb potential in the hydrogen atom, the orbital angular momentum commutes with the Hamiltonian, allowing simultaneous eigenstates of the energy and angular momentum operators. The magnitude of the orbital is specified by the orbital quantum number l, a non-negative (l = 0, 1, 2, \dots), while its projection along the z-axis is given by the m_l, which takes $2l + 1 values ranging from -l to +l. These quantum numbers emerge from the commutation relations of the components, [\mathbf{L}_i, \mathbf{L}_j] = i\hbar \epsilon_{ijk} \mathbf{L}_k, which impose constraints and limit the possible states. The eigenvalue equations for the squared magnitude and z-component are: \mathbf{L}^2 |l, m_l \rangle = \hbar^2 l(l+1) |l, m_l \rangle L_z |l, m_l \rangle = \hbar m_l |l, m_l \rangle where |l, m_l \rangle denotes the common eigenstates. Notably, the eigenvalue for \mathbf{L}^2 is \hbar^2 l(l+1) rather than \hbar^2 l^2, reflecting the quantum correction to the classical \mathbf{L}^2 = l^2 \hbar^2. In the position basis, the angular part of the wave function consists of spherical harmonics Y_{l m_l}(\theta, \phi), which serve as the normalized eigenfunctions of \mathbf{L}^2 and L_z on the unit sphere: Y_{l m_l}(\theta, \phi) = (-1)^{m_l} \sqrt{\frac{(2l+1)(l - m_l)!}{4\pi (l + m_l)!}} P_l^{m_l}(\cos \theta) e^{i m_l \phi} for m_l \geq 0, with Y_{l, -m_l} = (-1)^{m_l} Y_{l m_l}^* (associated Legendre functions P_l^{m_l}). These functions form an orthonormal basis for expanding angular dependencies in central-force problems, ensuring the separability of the into radial and angular parts. This framework was established in the through Erwin Schrödinger's solution to the via the time-independent , where the orbital quantization naturally arose from the boundary conditions on the spherical domain. Schrödinger's work built on earlier semiclassical models, such as Bohr's, but provided a fully wave-mechanical description, confirming the nature of l and introducing the l(l+1) eigenvalue structure.

Spin Angular Momentum

Spin angular momentum, often denoted as S, represents an intrinsic property of elementary particles, arising independently of any spatial rotation or orbital motion. Unlike orbital angular momentum, which stems from a particle's and in space, spin is a fundamental characteristic akin to mass or charge, manifesting as a form of internal . This property was first conceptualized to account for observed spectral anomalies in . The quantum description of spin involves two key quantum numbers: the spin quantum number s, which determines the magnitude of the spin angular momentum, and the magnetic quantum number m_s, which specifies its projection along a chosen axis, typically the z-axis. The value of s can be either an integer (0, 1, 2, ...) for bosons or a half-integer (1/2, 3/2, ...) for fermions, reflecting the particle's statistics. The projection m_s takes discrete values from -s to +s in integer steps of 1. For instance, particles with s = 1/2, such as electrons, have m_s = \pm 1/2. These quantum numbers fully characterize the spin state in the absence of external fields. The operators associated with spin angular momentum satisfy eigenvalue equations that quantify its measurable properties. The square of the total spin operator has eigenvalues given by S^2 \, |s, m_s\rangle = \hbar^2 s(s+1) \, |s, m_s\rangle, indicating that the magnitude of the spin angular momentum is \sqrt{s(s+1)} \, \hbar, not simply s \hbar. The z-component operator S_z yields S_z \, |s, m_s\rangle = \hbar m_s \, |s, m_s\rangle, with |s, m_s⟩ denoting the common eigenstates. These relations highlight the quantized nature of spin, where measurements yield discrete outcomes. For the simplest case of s = 1/2, such as the , the spin operators are represented using the Pauli spin matrices σ_x, σ_y, and σ_z, with the spin operators defined as S_i = (\hbar/2) σ_i for i = x, y, z. The matrices 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}. These matrices satisfy the commutation relations of operators and form the basis for describing -1/2 systems in . The concept of electron spin was proposed in 1925 by and to explain the splitting in atomic spectra, attributing it to an intrinsic of s = 1/2 for the . This proposal resolved discrepancies in the Sommerfeld model by introducing a associated with the spin. Earlier, the 1922 Stern-Gerlach experiment by and demonstrated the quantization of projections in silver atoms, providing experimental evidence for discrete m values, later interpreted in the context of spin. Representative examples illustrate the diversity of spin values. The , a fundamental , has s = 1/2, as confirmed by its spectroscopic behavior and Dirac's relativistic . The proton, a composite , also possesses s = 1/2, evident from and hyperfine splitting in . In contrast, the , the massless mediating , has s = 1, reflecting its polarization states.

Vector Coupling of Angular Momenta

Addition of Two Angular Momenta

In , the total of a composite system consisting of two angular momenta \mathbf{J_1} and \mathbf{J_2} is defined by the operator \mathbf{J} = \mathbf{J_1} + \mathbf{J_2}, where \mathbf{J_1} acts on the of the first subsystem and \mathbf{J_2} on the second, with the total space being the . Although the components of \mathbf{J_1} and \mathbf{J_2} do not commute across subsystems, the squared magnitude J^2 and the z-component J_z commute with each other and with the individual J_1^2, J_1_z, J_2^2, and J_2_z. The total angular momentum operators satisfy the standard commutation relations for angular momentum: [J_x, J_y] = i \hbar J_z, \quad [J_y, J_z] = i \hbar J_x, \quad [J_z, J_x] = i \hbar J_y. These relations follow directly from the corresponding commutation rules for \mathbf{J_1} and \mathbf{J_2}, since cross-subsystem commutators vanish. To facilitate the analysis of eigenvalues, ladder operators are introduced as J_\pm = J_x \pm i J_y, which raise or lower the z-projection eigenvalue m_j while acting within the coupled basis. Specifically, J_\pm |j, m_j\rangle = \hbar \sqrt{j(j+1) - m_j(m_j \pm 1)} \, |j, m_j \pm 1\rangle, preserving the total j value. The possible eigenvalues of J^2 are \hbar^2 j(j+1), where j ranges from |j_1 - j_2| to j_1 + j_2 in steps, and for each j, m_j ranges from -j to +j in or steps matching the subsystems. This range arises from the conservation of the z-projection, m_j = m_1 + m_2, and the in the model: the magnitude of \mathbf{J} cannot exceed j_1 + j_2 or fall below |j_1 - j_2|. The proceeds by considering the highest possible m_j = j_1 + j_2, which defines the maximum j = j_1 + j_2; applying the lowering J_- repeatedly generates a chain of states until m_j = -(j_1 + j_2), but to states with lower total m requires additional j multiplets down to |j_1 - j_2| to account for the full dimensionality of the space, (2j_1 + 1)(2j_2 + 1). A representative case is the of orbital \mathbf{L} (with l) and \mathbf{S} (with s = 1/2) for a single , yielding total j = l + 1/2 or j = l - 1/2 (except for l = 0, where j = 1/2). For multi-particle systems, the total is obtained by successive pairwise : first combine two particles to form an intermediate \mathbf{J_{12}}, then add the third as \mathbf{J} = \mathbf{J_{12}} + \mathbf{J_3}, and so on, with the allowed j values determined iteratively at each step to ensure conservation of total m_j and the overall dimension. This scheme generalizes the two-body while respecting the algebraic structure of the rotation group.

Clebsch-Gordan Coefficients

Clebsch-Gordan coefficients, denoted as \langle j_1 m_1 j_2 m_2 | J M \rangle, serve as the expansion coefficients that express the coupled total eigenstates in the uncoupled basis. Specifically, the total state |J M\rangle is given by |J M\rangle = \sum_{m_1 m_2} \langle j_1 m_1 j_2 m_2 | J M \rangle |j_1 m_1\rangle |j_2 m_2\rangle, where the sum runs over all m_1 and m_2 satisfying m_1 + m_2 = M, and the coefficients vanish otherwise. These coefficients arise naturally when coupling two angular momenta j_1 and j_2 to form a total J, ensuring the states transform irreducibly under rotations. The Clebsch-Gordan coefficients possess several key properties that facilitate their use in quantum mechanical calculations. They are real numbers, satisfying \langle j_1 m_1 j_2 m_2 | J M \rangle = \langle J M | j_1 m_1 j_2 m_2 \rangle. relations hold, such as \sum_{m_1 m_2} \langle j_1 m_1 j_2 m_2 | J M \rangle \langle j_1 m_1' j_2 m_2' | J M \rangle^* = \delta_{m_1 m_1'} \delta_{m_2 m_2'}, and ensures the between bases is unitary. Phase conventions are crucial for uniqueness; the widely adopted Condon-Shortley convention stipulates that the coefficient \langle j_1 j_1 j_2 M - j_1 | J M \rangle is real and positive when M = J, introducing a factor of (-1)^{j_1 - m_1} in certain definitions to align with . Analytical expressions for Clebsch-Gordan coefficients exist for simple cases, often derived via recursion relations from ladder operators. For coupling orbital angular momentum l with s = 1/2, the coefficients for j = l + 1/2 are \langle l, m_j - 1/2, 1/2, 1/2 | j m_j \rangle = \sqrt{\frac{l + m_j + 1/2}{2l + 1}}, \quad \langle l, m_j + 1/2, 1/2, -1/2 | j m_j \rangle = \sqrt{\frac{l - m_j + 1/2}{2l + 1}}, with the states normalized under the Condon-Shortley phase. For j = l - 1/2, the coefficients are obtained by orthogonal combinations, such as \langle l, m_j - 1/2, 1/2, 1/2 | j m_j \rangle = -\sqrt{\frac{l - m_j + 1/2}{2l + 1}}, \quad \langle l, m_j + 1/2, 1/2, -1/2 | j m_j \rangle = \sqrt{\frac{l + m_j + 1/2}{2l + 1}}. These formulas follow from applying the lowering operator J_- to the highest-weight state and ensuring normalization. Representative examples illustrate these coefficients for low values, such as a p-electron with l=1, s=1/2. The possible total j = 3/2 or $1/2. For j=3/2: | m_j | \langle 1, m_l, 1/2, m_s | 3/2, m_j \rangle | |---------|------------------------------------------------| | 3/2 | \langle 1 1, 1/2 1/2 | 3/2 3/2 \rangle = 1 | | 1/2 | \langle 1 0, 1/2 1/2 | 3/2 1/2 \rangle = \sqrt{2/3}, \langle 1 1, 1/2 -1/2 | 3/2 1/2 \rangle = \sqrt{1/3} | | -1/2 | \langle 1 -1, 1/2 1/2 | 3/2 -1/2 \rangle = \sqrt{1/3}, \langle 1 0, 1/2 -1/2 | 3/2 -1/2 \rangle = \sqrt{2/3} | | -3/2 | \langle 1 -1, 1/2 -1/2 | 3/2 -3/2 \rangle = 1 | For j=1/2, m_j=1/2: \langle 1 0, 1/2 1/2 | 1/2 1/2 \rangle = -\sqrt{1/3}, \langle 1 1, 1/2 -1/2 | 1/2 1/2 \rangle = \sqrt{2/3}, with analogous forms for m_j=-1/2. These values satisfy and are computed using the analytical expressions above. Historically, Clebsch-Gordan coefficients originated in 19th-century classical , developed by Alfred Clebsch and Paul Gordan for decomposing tensor products of binary forms. Their quantum mechanical adaptation for was introduced by Eugene P. Wigner in 1931. For computational purposes beyond , Racah coefficients facilitate recoupling of multiple angular momenta, transforming between schemes like ((j_1 j_2) j_{12} j_3) J and (j_1 (j_2 j_3) j_{23}) J. Defined as W(j_1 j_2 J j_3; j_{12} j_{23}), these are related to 6j symbols and enable efficient evaluation of matrix elements in complex systems via or hypergeometric series.

Properties of Total Angular Momentum

Definition and Magnitude

In quantum mechanics, the total angular momentum \vec{J} of a single particle, such as an electron in an atom, is defined as the vector sum \vec{J} = \vec{L} + \vec{S}, where \vec{L} is the orbital angular momentum and \vec{S} is the spin angular momentum. The square of this total angular momentum operator, J^2, has eigenvalues \hbar^2 j(j+1), where j is the total angular momentum quantum number, a non-negative value that characterizes the magnitude associated with the corresponding eigenstates. The magnitude of the total angular momentum vector is thus |\vec{J}| = \sqrt{j(j+1)} \hbar, rather than simply j \hbar, reflecting the inherent quantum in the simultaneous of all components of \vec{J}. This contrasts with classical addition, where the total angular momentum would have a definite length and direction; in , the \vec{J} precesses around the quantization axis, such as the z-axis, due to the commutation relations among its components. Notation conventions distinguish between single-particle and multi-particle systems: lowercase j denotes the quantum number for a single particle, while uppercase J is used for the total angular momentum of a composite system, such as an entire . In the relativistic framework, the incorporates spin-orbit coupling naturally, conserving the total angular momentum \vec{J} as a good while treating orbital and contributions on equal footing.

Possible Values and Projections

In , the total angular momentum quantum number j for a system combining orbital l and angular momentum s takes values ranging from |l - s| to l + s in integer steps. For example, in the case of an with s = 1/2, the possible j values are l + 1/2 and l - 1/2. For a fixed j, the quantum number m_j along the z-axis takes $2j + 1 equally spaced values from -j to +j in steps of 1, corresponding to the degeneracy of the state in the absence of an external field. These projections arise from the eigenvalues of the J_z, with the total number of states for a given j reflecting the in three dimensions. In electric dipole transitions, selection rules govern changes in j and m_j: \Delta j = 0, \pm 1 (excluding $0 \to 0 transitions) and \Delta m_j = 0, \pm 1, depending on the light polarization. These rules ensure conservation of during emission or , where the 's of 1 limits the possible changes. The introduces a that lifts the m_j degeneracy, splitting the energy levels by \Delta E = g \mu_B B m_j, where g is the given by g = 1 + \frac{j(j+1) + s(s+1) - l(l+1)}{2j(j+1)}, \mu_B is the , and B is the field strength. This splitting reveals the magnetic properties tied to the total orientation. due to spin-orbit coupling further splits levels within a given l and s, with the energy shift proportional to \frac{1}{2} [j(j+1) - l(l+1) - s(s+1)], leading to distinct energies for different j values. The magnitude of the total is \sqrt{j(j+1)} \hbar, but the discrete j and m_j values determine the observable projections and transitions.

Applications in Physics

Atomic and Molecular Spectroscopy

In , the total angular momentum quantum number j is essential for understanding the arising from -orbit , which interacts the orbital \mathbf{L} and spin angular momentum \mathbf{S} of electrons to form the total \mathbf{j} = \mathbf{L} + \mathbf{S}. This coupling splits degenerate energy levels into components characterized by different j values, with the magnitude of splitting proportional to the expectation value of \mathbf{L} \cdot \mathbf{S}. In the Russell-Saunders () coupling approximation, valid for light atoms where spin-spin and orbit-orbit interactions are weaker than spin-orbit coupling, atomic states are described by term symbols of the form ^{2S+1}L_j, where $2S+1 is the multiplicity, L is the total orbital angular momentum quantum number (denoted by letters S for 0, P for 1, for 2, etc.), and the subscript j specifies the total angular momentum quantum number ranging from |L - S| to L + S. The energy shift for each j is given by \Delta E \propto j(j+1) - L(L+1) - S(S+1), leading to observable splittings in spectral lines. A classic illustration is the sodium D-line doublet in the , produced by transitions from the to the 3s . The is ^2S_{1/2} with L=0, S=1/2, so j=1/2. The 3p state, with L=1, S=1/2, splits into ^2P_{3/2} (j=3/2) and ^2P_{1/2} (j=1/2) due to spin-orbit , with the j=3/2 level higher in for inverted multiplets in atoms. The transitions ^2P_{3/2} \to ^2S_{1/2} (D2 line) and ^2P_{1/2} \to ^2S_{1/2} (D1 line) occur at vacuum wavelengths of 588.9950 nm and 589.5924 nm, respectively, with an intensity ratio of approximately 2:1 reflecting the degeneracy of the upper levels. These lines, first resolved in 1890, provided early confirmation of the predicted by the . Hyperfine structure introduces additional splitting in atomic spectra due to the magnetic interaction between the total electronic angular momentum \mathbf{j} and the nuclear angular momentum \mathbf{I}, forming the total angular momentum \mathbf{F} = \mathbf{j} + \mathbf{I}. The possible F values range from |j - I| to j + I in steps, with the hyperfine shift typically \Delta E \propto F(F+1) - j(j+1) - I(I+1), much smaller than but resolvable in high-resolution . For sodium (I=3/2), the D2 line's ^2P_{3/2} level (j=3/2) splits into F=1,2 components, while the splits into F=1,2, resulting in multiple hyperfine transitions observable as narrow lines within the fine structure doublet. This structure has been precisely measured using laser , aiding applications in atomic clocks and precision tests of fundamental symmetries. In molecular , the total quantum number is adapted to the cylindrical of diatomic molecules through Hund's coupling cases, which describe different limits of depending on the relative strengths of spin-orbit, spin-rotation, and rotational interactions. Hund's case (a) applies to light molecules or states with significant spin-orbit coupling (\Lambda \neq 0, where \Lambda is the projection of L along the internuclear axis), where individual electronic orbital and spin projections \Lambda and \Sigma couple to form the total projection \Omega = |\Lambda + \Sigma|, and then the nuclear rotation N couples with \Omega to yield the total angular momentum J = N + \Omega (with J \geq \Omega). Term symbols are written as ^{2S+1}\Lambda_{\Omega} (with parity superscript), and rotational levels show Lambda-doubling for non-\Sigma states. Cases (b) and (c) handle weaker spin-orbit coupling, but case (a) dominates for many electronic transitions in molecules like O2 or NO, enabling interpretation of band spectra. Experimental spectra of alkali metals, such as the fine and hyperfine structure in sodium and potassium vapor, exhibit wavelength shifts that precisely match predictions from total angular momentum coupling. For instance, the sodium D-line separation of 0.597 nm corresponds to the spin-orbit splitting calculated from j values, with hyperfine components shifted by ~1.8 GHz in the ground state, as verified by interferometric and laser absorption measurements. Similar agreement in rubidium and cesium spectra has confirmed the Russell-Saunders scheme across the alkali series, providing benchmarks for quantum mechanical models of atomic interactions.

Nuclear and Particle Physics

In the , nucleons occupy discrete energy levels characterized by orbital quantum number l and intrinsic s = 1/2, leading to a total j = l \pm 1/2 for each subshell due to strong spin-orbit coupling. Subshells are filled according to the , with examples including the p_{3/2} (where j = 3/2) and p_{1/2} (where j = 1/2) levels, which accommodate up to $2j + 1 nucleons each. This structure arises from an average potential resembling a modified by spin-orbit interactions, explaining the stability of nuclei with specific numbers. Closed subshells in this model produce —2, 8, 20, 28, 50, 82, and 126—corresponding to filled j-subshells where the total angular momentum is zero, resulting in enhanced and nuclear stability due to the $2j + 1 degeneracy of each level. For instance, the filling of the $1p_{3/2} and $1p_{1/2} subshells contributes to the magic number 8. In , the total angular momentum quantum number J classifies , with baryons exhibiting ground states of J = 1/2 (e.g., nucleons like the proton and ) or J = 3/2 (e.g., the Δ resonances), determined by the vector coupling of constituent spins and orbital contributions. Regge trajectories describe excited states by plotting J linearly against the square of the mass m^2, revealing nearly universal slopes indicative of underlying string-like dynamics in . For example, the ρ meson trajectory shows J increasing with m^2, connecting vector mesons (J = 1) to higher-spin tensors. Conservation of total J governs particle decays and reactions, imposing selection rules alongside parity considerations; transitions must satisfy |\Delta J| \leq 1 for electromagnetic decays or stricter limits in weak processes, forbidding certain modes like the direct $0^+ \to 0^+ transition without . In the , the total J of hadrons emerges from combining quark spins (s_q = 1/2) and orbital angular momenta, yielding J = 1/2 for the octet (symmetric spin-flavor wave functions) or J = 3/2 for the decuplet (fully symmetric). A representative example is the deuteron, the bound np state with total J = 1, primarily from l = 0, s = 1 (S-wave, ~96% probability) admixed with a small l = 2, s = 1 (D-wave) component to account for its quadrupole moment.

References

  1. [1]
    8.9: The Allowed Values of J - the Total Angular Momentum Quantum Number
    ### Definition and Key Properties of J (Total Angular Momentum Quantum Number)
  2. [2]
    Total Angular Momentum - an overview | ScienceDirect Topics
    Total angular momentum is defined as a quantum property of an atom characterized by the total angular momentum quantum number \( J \), which contributes to the ...
  3. [3]
    [PDF] Derivation of Angular Momentum Rules in Quantum Mechanics
    Mar 23, 2015 · In a composite system, the total angular momentum is computed as a vector sum: LT = Pi Li. We would like to determine the stationary states of ...
  4. [4]
    [PDF] Quantum Mechanics - Richard Fitzpatrick
    Finally, in Chapters 10 and 11, we shall examine spin angular momentum, and the addition of orbital and spin angular momentum, respectively. The second part ...
  5. [5]
    [PDF] Quantisierung als Eigenwertproblem
    1926. Page 25. Quantisierung als Eigenwertproblem. 133 und Radialquant, was heute allgemein als korrekt angesehen wird; allein es fehlt noch die zur ...
  6. [6]
    [PDF] 03 - Spin Angular Momentum
    ▻ Charge: e = −1.60217662 10−19 C. ▷ However, an electron also possesses an intrinsic (i.e., non-orbital) angular momentum known as spin.<|control11|><|separator|>
  7. [7]
    30.8 Quantum Numbers and Rules – College Physics
    Intrinsic Spin Angular Momentum Is Quantized in Magnitude and Direction ... S = s ( s + 1 ) h 2 π ( s = 1 / 2 for electrons ) ,. where s is defined to be the spin ...
  8. [8]
    [PDF] ADDITION OF ANGULAR MOMENTUM - MIT OpenCourseWare
    Dec 12, 2013 · It is important to note that had we added the two angular momenta with some arbitrary coefficients we would not have got an angular momentum.
  9. [9]
    [PDF] 44. Clebsch-Gordan coefficients
    Note: A square-root sign is to be understood over every coefficient, e.g., for −8/15 read −p8/15. Y 0. 1 = r. 3. 4π cosθ. Y 1. 1 = −r.
  10. [10]
    Clebsch-Gordan Coefficients - SymPy 1.14.0 documentation
    Clebsch-Gordan coefficients describe the angular momentum coupling between two systems. The coefficients give the expansion of a coupled total angular momentum ...
  11. [11]
    Properties of Clebsch–Gordan numbers - IOP Science
    Clebsch-Gordan numbers are the multiplicity numbers for the occurrence of the states of total angular momentum in the coupling of n angular momenta. These ...
  12. [12]
    [PDF] Clebsch-Gordan coefficients and the tensor spherical harmonics
    The general expressions for the Clebsch-Gordan coefficients in terms of j, mℓ, ℓ, s and ms are very complicated to write down. Nevertheless, the explicit ...
  13. [13]
    Clebsch-Gordan Coefficient -- from Wolfram MathWorld
    Clebsch-Gordan coefficients are mathematical symbol used to integrate products of three spherical harmonics. Clebsch-Gordan coefficients commonly arise in ...
  14. [14]
    [PDF] The Clebsch-Gordan Coefficients - SciSpace
    Wigner, Gruppentheorie und ihre Anwendung auf die Quantenmechanik der Atomspektren, Braun- schweig (1931). Downloaded from https://academic.oup.com/ptp ...
  15. [15]
    [PDF] angular - - ladder
    Feb 13, 2020 · There are several angular momentum operators: total angular momentum (usually denoted J), orbital angular momentum (usually denoted L), and spin ...
  16. [16]
    The Feynman Lectures on Physics Vol. III Ch. 18: Angular Momentum
    In the last chapter we developed the idea of the conservation of angular momentum in quantum mechanics, and showed how it might be used to predict the ...
  17. [17]
    [PDF] Phys 487 Discussion 5 – Atomic Structure - Course Websites
    As in classical mechanics, we use lower-case letters to refer to single-particle properties and CAPITAL letters to refer to TOTAL properties of the system. For ...
  18. [18]
    The Dirac Equation
    The Dirac equation naturally conserves total angular momentum but not the orbital or spin parts of it. We can also see that the helicity, or spin along the ...
  19. [19]
    [PDF] Total Angular momentum - SIUC Physics WWW2
    Even though quantum mechanically j can take any value between l + s to |l − s| (implying a number of degenerate states), atoms usually pick either the highest ...
  20. [20]
    [PDF] 04 - Addition of Angular Momentum
    ▷ Saw that when spin l is added to spin 1/2 then possible values of total angular momentum quantum number are j = l ± 1/2. ▷ By analogy, when spin 1/2 is ...
  21. [21]
    [PDF] chapter 12 total angular momentum
    Total angular momentum is the sum of orbital and spin angular momenta, ~ J = ~ L + ~ S, where ~L and ~S are orbital and spin angular momenta.
  22. [22]
    [PDF] Energy Levels in Atoms 2018 - UCI Chemistry
    J = L + S is the total angular momentum of the atom. Possible quantum numbers for J: J = (L + S), (L + S – 1), (L + S – 2), ... | L-S |. The number of J ...<|control11|><|separator|>
  23. [23]
    [PDF] UCI PHYSICS / CHEM 207 - Applied Physical Chemistry
    Jul 27, 2022 · (Angular Momentum) Atomic Selection "rules". Orbital angular momentum ... Total angular momentum: 𝚫J = 0, ±1 … and 𝚫S = 0… and ...
  24. [24]
    [PDF] LECTURE NOTES ON QUANTUM MECHANICS
    ... the selection rules associated with atomic or nu- clear transitions. In the dipole approximation the transition operator behaves as if it has J = 1. Thus ...
  25. [25]
    [PDF] THE ZEEMAN EFFECT - Rutgers Physics
    where gL is called the Lande g-factor and is given by. gL =1+. J(J +1)+ S(S +1) ... correspond to the total angular momentum of all the electrons in the atom.
  26. [26]
    [PDF] The calculation of atomic and molecular spin-orbit coupling matrix ...
    2[J(J + 1) − L(L + 1) − S(S + 1)]. This equals A/2 for J = 3/2 and −A for ... S = MS ∓ 1. Note that we have designated the spin-orbit coupling ...
  27. [27]
    [PDF] Sodium D Line Data - Daniel A. Steck
    May 27, 2000 · For the ground state in sodium, L = 0 and S = 1/2, so J = 1/2; for the first excited state, L = 1, so J = 1/2 or J = 3/2. The energy of any ...<|control11|><|separator|>
  28. [28]
    NIST: Atomic Spectra Database Lines Form
    NIST Atomic Spectra Database Lines Form. Main Parameters: Spectrum eg, Fe I or Na;Mg; Al or mg i-iii or 198Hg I. Search for Wavelength, Wavenumber, Photon, ...
  29. [29]
    NIST: Diatomic Molecules - 2-Pi-Ground State Molecules
    Hund's coupling case (a) is employed as a starting point. The rotational levels are defined with the quantum number Ω, the absolute value of the projection of ...Missing: spectroscopy | Show results with:spectroscopy
  30. [30]
    [PDF] ELEMENTARY THEORY OF NUCLEAR SHELL STRUCTURE
    MARIA GOEPPERT MAYER. Advisory Editor. ELEMENTARY THEORY OF. NUCLEAR SHELL STRUCTURE. Page 2. Page 3. Page 4. ELEMENTARY THEORY OF. NUCLEAR SHELL STRUCTURE.
  31. [31]
    [PDF] Mayer–Jensen Shell Model and Magic Numbers
    orbital angular momentum have two possible energies, depending on whether their spin is parallel or antiparallel to the orbital motion. This splitting of ...Missing: jls | Show results with:jls
  32. [32]
    Shell model
    The main ingredients for the success of the nuclear shell model are the idea of independent-particle motion in nuclei, and strong spin-orbit coupling.Missing: jls | Show results with:jls
  33. [33]
    [PDF] 15. Quark Model - Particle Data Group
    May 31, 2024 · The meson spin J is given by the usual relation |` − s| ≤ J ≤ |` + s|, where s = 0 (antiparallel quark spins) or s = 1 (parallel quark spins).
  34. [34]
    Regularities in hadron systematics, Regge trajectories and a string ...
    Jan 8, 2007 · For convenience of a linear-regression analysis, we prefer to express M 2 as a function of J since the angular momentum does not have ...
  35. [35]
    [PDF] slac-pub-17201.pdf
    ... Regge trajectories are nearly identical when one plots M. 2 versus the hadron angular momentum J. For example, the spectroscopy of q¯q mesons are well ...
  36. [36]
    [PDF] March 28, 2015) Deuteron (nuclei of deuterium) - bingweb
    Mar 28, 2015 · Spin angular momentum S = 1 (in the units of ħ ). Orbital angular momentum L = 0. The total angular momentum: J = 1,. The intrinsic parity of ...