Fact-checked by Grok 2 weeks ago

Selection rule

In , a selection rule is a that determines whether a transition between two quantum states—such as energy levels in an or —is allowed or forbidden, primarily in the context of interactions with like or of . These rules specify the permissible changes in quantum numbers, such as or , thereby dictating the probability of a transition occurring; allowed transitions have high rates (often ~10^8 s^{-1}), while forbidden ones are suppressed or negligible under typical conditions. Selection rules originate from fundamental symmetries of the quantum system, including conservation of parity, , and , as well as the tensorial nature of the interaction operator (e.g., electric dipole for common radiative transitions). In , they are essential for interpreting spectra, as they filter observable lines and reveal underlying electronic configurations; for example, forbidden transitions may become weakly visible in low-density environments like stellar atmospheres. Key examples include the , which requires a change in orbital parity (Δl = ±1 for electric dipole transitions, prohibiting s-to-s or p-to-p shifts), and the spin selection rule (ΔS = 0), which conserves total spin multiplicity and forbids singlet-to-triplet transitions in most cases. For total angular momentum J in atoms, ΔJ = 0, ±1 (with J=0 to J=0 forbidden), ensuring compliance with angular momentum transfer. These principles extend to molecular , , and particle decays, providing a unified framework for predicting transition intensities across quantum systems.

Fundamentals

Definition

Selection rules in are restrictions that govern the allowed transitions between quantum states of a , determining whether such a change is probable and observable (allowed) or highly improbable and effectively negligible (forbidden). These rules arise from the fundamental symmetries and conservation laws of , ensuring that only certain transitions couple effectively to external perturbations like electromagnetic fields. For instance, the conservation of serves as a key underlying many selection rules, limiting the possible changes in angular momentum quantum numbers during a transition. The concept originated in the early development of during the 1920s, with foundational contributions from and contemporaries who sought to reconcile atomic spectra with the emerging formalism of wave mechanics and . Dirac's work in particular formalized these restrictions through the quantum theory of radiation, where transitions are analyzed via the interaction describing light-matter coupling. By 1926–1927, initial justifications for selection rules had appeared in the new quantum framework, building on empirical observations from that not all theoretically possible transitions occur with equal intensity. In general, a transition is allowed if the , given by the integral \langle \psi_f | \hat{\mu} | \psi_i \rangle = \int \psi_f^* \hat{\mu} \psi_i \, d\tau \neq 0, where \psi_i and \psi_f are the initial and final wavefunctions, respectively, and \hat{\mu} is the , is nonzero; otherwise, the transition is forbidden in the lowest-order approximation. Forbidden transitions can still occur through higher-order processes, such as or interactions, but these are significantly weaker, often by factors of $10^{-3} to $10^{-6} compared to allowed electric dipole transitions. This binary classification—allowed versus forbidden—provides a practical framework for predicting intensities and interpreting experimental observations in atomic and molecular systems.

Quantum Mechanical Basis

In , selection rules governing transitions between discrete energy eigenstates originate from applied to systems interacting with external fields, such as . The is decomposed as H = H_0 + H'(t), where H_0 describes the unperturbed system and H'(t) is the time-dependent . Within first-order , the transition probability from an initial state |i\rangle to a final state |f\rangle is determined by the matrix element \langle f | H' | i \rangle. provides the transition rate w_{i \to f} = \frac{2\pi}{\hbar} |\langle f | H' | i \rangle|^2 \rho(E_f), where \rho(E_f) is the density of final states, assuming a weak, H'(t) = H' e^{-i\omega t} + \mathrm{H.c.} with E_f = E_i \pm \hbar \omega. This framework implies that transitions are allowed only if the matrix element is non-vanishing, establishing the foundational criterion for selection rules in processes like or . For electromagnetic interactions, the relevant perturbation in the electric dipole approximation—valid when the field wavelength exceeds the system's spatial extent—is H' = -\boldsymbol{\mu} \cdot \mathbf{E}, with the dipole operator \boldsymbol{\mu} = -e \mathbf{r} for a single (or summed over charges for multi-particle systems). The transition rate then depends on |\langle f | \boldsymbol{\mu} | i \rangle|^2, and selection rules arise directly from the properties of the wavefunctions \psi_i and \psi_f that make this non-zero. Specifically, the operator \boldsymbol{\mu} transforms as a under spatial rotations and reflections, restricting allowed transitions to those where the initial and final states couple through this vectorial form; scalar operators would enforce different constraints. This derivation highlights how the perturbation's operator form dictates the quantum mechanical conditions for observable transitions without requiring full computation of the eigenstates. Symmetry analysis via further refines these rules by classifying states and operators according to irreducible representations (irreps) of the system's , such as the group SO(3) or point groups for molecules. The matrix element \langle f | O | i \rangle vanishes unless the irrep of the operator O (e.g., the components) appears in the decomposition of the irrep of |f\rangle with the complex conjugate of the irrep of |i\rangle; otherwise, the over symmetric coordinates yields zero by . This character-based criterion, rooted in , provides a systematic way to predict allowed transitions across diverse systems. These selection rules hold rigorously only under idealized assumptions, such as isolated atoms or molecules with exact symmetries and neglect of higher multipole terms in the interaction. In real systems, perturbations like spin-orbit coupling, magnetic fields, or vibrational modes mix states, enabling weakly allowed "forbidden" transitions through higher-order processes in , where the rate is suppressed by factors like (V/\Delta E)^2 relative to allowed ones, with V the mixing potential and \Delta E the energy separation./12%3A_Time-Dependent_Perturbation_Theory/12.13%3A_Forbidden_Transitions)

Importance in Spectroscopy

Selection rules play a central role in by determining which quantum transitions between energy levels are permitted, thereby predicting the positions and relative strengths of spectral lines without requiring exhaustive quantum mechanical computations for every atomic or molecular system. This predictive capability stems from conservation laws, such as those governing and , which restrict transitions to those compatible with the photon's properties during or . By identifying allowed transitions a priori, spectroscopists can efficiently assign observed features in spectra to specific physical processes, streamlining the interpretation of experimental data across diverse systems. The implications for spectral intensities are equally significant: allowed transitions dominate the observed spectra with strong signals, whereas forbidden transitions manifest as weak lines, often arising from perturbations like vibronic coupling or external fields. This distinction enables detailed analysis of subtle spectral perturbations, revealing information about environmental influences or higher-order effects in the system. For instance, selection rules, such as those limiting the change in total to ΔJ = 0, ±1, provide a primary for anticipating these intensity patterns in both and molecular spectra. In practical applications, selection rules are indispensable for experimental techniques like , where they guide the choice of wavelengths to excite specific transitions, enabling precise control over atomic and . In , they facilitate the of stellar atmospheres and nebulae by predicting observable lines, such as those in spectra, to infer elemental compositions, temperatures, and densities from remote observations. Similarly, in chemical , selection rules in and allow the identification of molecular structures and bonding through allowed vibrational and rotational transitions. However, these rules have limitations in extreme conditions, such as intense electromagnetic fields, where higher-order interactions can enable nominally forbidden transitions through processes like multi-photon absorption or coherent effects.

Angular Momentum Selection Rules

Total Angular Momentum Change (ΔJ)

In electric dipole transitions, the selection rule for the total angular momentum quantum number J stipulates that the change \Delta J = J_f - J_i must satisfy \Delta J = 0, \pm 1, with the exception that transitions between states with J = 0 and J' = 0 are forbidden. This rule arises because the electric dipole operator is a rank-1 tensor, limiting the possible angular momentum transfers to those compatible with coupling to a photon of spin 1. The physical basis for this selection rule stems from the conservation of total in the between the or molecular system and the . The emitted or absorbed carries an intrinsic angular momentum of \hbar (corresponding to spin 1), which can couple with the initial state's angular momentum \mathbf{J}_i to yield the final state's \mathbf{J}_f, resulting in possible changes of 0 or \pm 1 in the magnitude of J. Violations of this rule would imply non-conservation of , rendering such transitions impossible in the dipole approximation. Mathematically, the transition \langle J_f, M_f | \hat{D} | J_i, M_i \rangle, where \hat{D} is the , is evaluated using the Wigner-Eckart theorem. This decomposes the element into a reduced matrix element and a Clebsch-Gordan coefficient: \langle J_f M_f | T^{(1)}_q | J_i M_i \rangle = \langle J_i M_i 1 q | J_f M_f \rangle \langle J_f || T^{(1)} || J_i \rangle / \sqrt{2J_f + 1}. The Clebsch-Gordan coefficient \langle J_i M_i 1 q | J_f M_f \rangle vanishes unless the angular momenta satisfy the |J_f - J_i| \leq 1 \leq J_f + J_i, enforcing the \Delta J = 0, \pm 1 condition (and prohibiting J_i = J_f = 0 since no valid coupling exists for two spin-1 particles to total J=0). For weaker, higher-order transitions beyond the electric dipole approximation, the \Delta J selection rule is relaxed due to the higher rank of the corresponding tensor operators. (M1) transitions, mediated by the magnetic dipole moment (also a rank-1 tensor), obey the same \Delta J = 0, \pm 1 (except $0 \leftrightarrow 0) but are distinguished by conservation rather than change. (E2) transitions, involving a rank-2 tensor, allow \Delta J = 0, \pm 1, \pm 2 (except $0 \leftrightarrow 0, $0 \leftrightarrow 1, and half-integer cases like $1/2 \leftrightarrow 3/2), enabling larger transfers at the cost of much lower transition probabilities. These rules play a crucial role in determining allowed spectral lines in and molecular .

Projection Quantum Number Change (ΔM)

In angular momentum transitions, the selection rule for the projection quantum number M_J governs the allowed changes \Delta M_J based on the polarization of the interacting . For \pi-polarized light, with the electric field aligned parallel to the quantization (typically the z-), the rule permits only \Delta M_J = 0. Conversely, for \sigma^+ and \sigma^- circularly polarized light, the allowed changes are \Delta M_J = +1 and \Delta M_J = -1, respectively. These rules ensure conservation of projection during the interaction with the photon's component along the propagation direction. The geometric origin of these rules stems from the decomposition of the interaction into spherical tensor components. The \mathbf{r} transforms as a rank-1 tensor, and its interaction with the light's \mathbf{E} depends on the field's orientation relative to the quantization axis. Linear \pi polarization corresponds to the q=0 component, preserving the z-projection of , while circular \sigma^\pm polarizations correspond to q=\pm1 components, transferring \pm \hbar along the axis. This is formalized in the element, where the angular part involves the integral of three : \langle J' M' | Y_{1q} | J M \rangle \propto \int Y_{J' M'}^* Y_{1q} Y_{J M} \, d\Omega, which is non-zero only if q = M' - M = \Delta M_J = 0, \pm 1, as determined by the properties of the 3j symbols or Clebsch-Gordan coefficients in the Wigner-Eckart theorem. In polarization spectroscopy, measuring \Delta M_J via selective excitation with polarized light distinguishes transition types by isolating specific angular momentum projections. For instance, \pi polarization excites only \Delta M_J = 0 sublevels, revealing parallel transition strengths, while \sigma polarization accesses \Delta M_J = \pm 1, highlighting perpendicular or helical components. This approach enhances resolution of Zeeman substructure and atomic orientation without relying on magnetic fields. These projection rules build on the total angular momentum change \Delta J = 0, \pm 1 to provide a complete description of allowed electric dipole transitions.

Parity and Other Conservation Laws

In quantum mechanics, parity conservation imposes a fundamental selection rule on electromagnetic transitions between atomic or molecular states. The parity operator \hat{P} is defined such that \hat{P} \psi(\mathbf{r}) = \psi(-\mathbf{r}), where \psi(\mathbf{r}) is the wave function of the system, assigning an eigenvalue of +1 (even parity) or -1 (odd parity) to eigenstates based on their behavior under spatial inversion. For electric dipole (E1) transitions, the transition dipole moment operator \boldsymbol{\mu} has odd parity, leading to the selection rule \Delta P = -1, meaning the initial and final states must have opposite parities for the matrix element \langle f | \boldsymbol{\mu} | i \rangle to be nonzero; if the product of the parities P_f P_i = +1, the integral vanishes due to symmetry. In contrast, magnetic dipole (M1) transitions involve an even-parity operator, requiring \Delta P = +1 (same parity) for allowed transitions. This parity rule originates from the intrinsic properties of the and the of the . The in an E1 carries odd intrinsic (P_\gamma = -1), ensuring overall conservation when combined with the s, as the total of the initial must match that of the final plus the emitted . For M1 transitions, the 's effective is even (P_\gamma = +1), aligning with the operator's . These rules are derived within the long-wavelength approximation, where higher-order multipole effects are negligible. Beyond parity, other conservation laws further constrain transitions in atomic spectroscopy. Energy conservation requires the photon energy to match the difference between initial and final state energies, while charge conservation mandates that the total electric charge remains unchanged, typically holding automatically for intra-atomic transitions without ionization. Linear momentum conservation, in the dipole approximation, implies \Delta \mathbf{k} = 0 for the atomic center-of-mass wave vector, as the photon's momentum is small compared to the atomic scale, neglecting recoil effects. These laws, combined with angular momentum rules, fully determine allowed transitions. In weak interactions, such as , parity is not conserved, relaxing these selection rules and allowing otherwise forbidden processes. This was experimentally demonstrated in the 1956 experiment by and collaborators, who observed asymmetric electron emission in the of polarized ^{60}Co nuclei at low temperatures, confirming violation in the weak force. In , this non-conservation permits transitions that violate electromagnetic rules, though strong and electromagnetic interactions still obey conservation.

Atomic Selection Rules

Electric Dipole Transitions

Electric dipole transitions represent the primary mechanism for radiative decay in atoms, dominating the emission and absorption lines observed in the visible and spectra due to their inherently strength compared to higher-order multipole processes. These transitions occur when an changes its state under the influence of the component of the electromagnetic wave, leading to the emission or absorption of a with energy matching the difference between initial and final states. In atomic systems, E1 transitions are "allowed" only if specific quantum mechanical conditions are met, enforcing the conservation of and while prohibiting others through vanishing transition matrix elements. The key selection rules for E1 transitions in atoms include restrictions on the orbital angular momentum and of the electrons. For a single electron, the orbital quantum number must change by Δl = ±1, ensuring the transition connects states with different orbital character, such as s to p or p to d orbitals. In multi-electron atoms, under the LS coupling approximation, the total remains unchanged, ΔS = 0, reflecting the spin-conserving nature of the electric dipole operator which does not couple degrees of freedom. These rules, combined with the general constraints of ΔJ = 0, ±1 (excluding J=0 to J=0), determine whether a is permitted. Parity conservation imposes the for E1 transitions, requiring a change in the of the overall wavefunction, typically from gerade (g) to ungerade (u) or vice versa, as the dipole operator is odd under spatial inversion. This rule arises because the transition matrix element vanishes if the initial and final states have the same , prohibiting intra-configuration transitions like those within the same subshell. Violations of these selection rules result in "forbidden" transitions that are suppressed by orders of magnitude. The strength of allowed E1 transitions is characterized by the oscillator strength f, a that measures the transition's probability relative to a classical . It is given by f \propto |\langle f | \mathbf{r} | i \rangle|^2, where \langle f | \mathbf{r} | i \rangle is the electric between final (f) and initial (i) states. Selection rules enforce f = 0 for forbidden transitions, explaining their weakness, while allowed E1 lines exhibit oscillator strengths on the order of 0.1 to 1, enabling their dominance in atomic spectra. For example, the transition in (1s to 2p) has f \approx 0.416, illustrating a strong allowed E1 process.

LS Coupling Scheme

In the Russell-Saunders or LS coupling scheme, which applies to light atoms (low Z) where spin-orbit interactions are weaker than electron-electron repulsions, the individual orbital angular momenta \mathbf{l}_i of the electrons couple to a total orbital angular momentum \mathbf{L} = \sum \mathbf{l}_i, and the spins \mathbf{s}_i couple to a total spin \mathbf{S} = \sum \mathbf{s}_i. These then couple to form the total \mathbf{J} = \mathbf{L} + \mathbf{S}. The resulting atomic energy levels are characterized by term symbols of the form ^{2S+1}\mathrm{L}_J, where $2S+1 is the multiplicity, \mathrm{L} denotes the orbital angular momentum (S for L=0, P for L=1, D for L=2, etc.), and J is the . For electric dipole (E1) transitions in LS coupling, the primary selection rules are \Delta L = 0, \pm 1 (excluding $0 \leftrightarrow 0), \Delta S = 0, and \Delta J = 0, \pm 1 (excluding $0 \leftrightarrow 0), along with a change in parity. These rules ensure that transitions occur between terms of the same multiplicity (^{2S+1}) and typically involve a change in L by \pm 1 for the strongest allowed lines, reflecting the vector addition properties of angular momenta in the dipole operator. Hund's rules integrate with LS coupling to identify the ground state term from a given electron configuration, guiding which excitations are possible. The rules state that the lowest-energy term maximizes S (highest multiplicity to minimize electron repulsion via parallel spins), then maximizes L for that S (to maximize orbital moment and reduce Coulomb energy), and finally sets J = |L - S| for shells less than half full or J = L + S for more than half full (due to spin-orbit effects). For instance, the carbon ground configuration $1s^2 2s^2 2p^2 yields the ^3\mathrm{P}_0 term as lowest via these rules, from which allowed excitations to singlet or triplet P or D terms follow the \Delta S = 0 and \Delta L = \pm 1 constraints. Representative examples illustrate these rules in practice. In alkali atoms like sodium (ground state $3s\ ^2\mathrm{S}_{1/2}), allowed E1 transitions excite the valence electron to $3p\ ^2\mathrm{P}_{1/2,3/2} states, satisfying \Delta l = \pm 1 for the electron (equivalent to \Delta L = 1 in the one-valence-electron approximation) and \Delta S = 0, producing the prominent yellow D-line doublet at 589 nm. In helium, the ground state $1s^2\ ^1\mathrm{S}_0 (singlet, per Hund's maximum L=0 for equivalent electrons) allows transitions to $1s2p\ ^1\mathrm{P}^\circ_1 (same multiplicity, \Delta L = 1), but forbids intercombination lines to triplet states like $1s2p\ ^3\mathrm{P}^\circ_J due to \Delta S = 0, separating singlet and triplet manifolds.

jj Coupling Scheme

In the jj coupling scheme, the total angular momentum of each is formed by first coupling its individual orbital angular momentum \mathbf{l}_i and \mathbf{s}_i to yield \mathbf{j}_i = \mathbf{l}_i + \mathbf{s}_i, with j_i = l_i \pm 1/2. These individual \mathbf{j}_i are then coupled together to produce the total atomic \mathbf{J} = \sum \mathbf{j}_i. This approach is essential for heavy atoms, where the spin-orbit interaction dominates over interelectronic repulsion, particularly for elements with Z > 30, such as (Z = 79). The spin-orbit coupling strength scales proportionally to Z^4 / n^3, where n is the principal quantum number, making relativistic effects pronounced in inner shells of these atoms. The scheme originates from the relativistic treatment in the , which describes as having intrinsic spin-orbit coupling without needing perturbative additions. For electric dipole transitions in the jj coupling regime, the selection rules require the change in the to satisfy \Delta J = 0, \pm 1 (excluding J = 0 to J = 0), and for the transitioning , \Delta j = 0, \pm 1 (excluding j = 0 to j = 0). Unlike lighter atoms where LS coupling allows rules based on \Delta L and \Delta S, the jj scheme lacks simple \Delta L or \Delta S restrictions because total L and S are not conserved quantum numbers. These rules also apply to the projection quantum numbers, with \Delta M_J = 0, \pm 1. In spectra of heavy atoms, jj coupling governs K-shell transitions, where an from the L shell ($2p, with j = 1/2 or $3/2) fills the K-shell vacancy ($1s, j = 1/2), adhering to \Delta j = \pm 1. This results in distinct lines like K\alpha_1 (from $2p_{3/2}) and K\alpha_2 (from $2p_{1/2}), observed in elements like . In contrast to the LS coupling scheme for lighter atoms, jj coupling better captures the relativistic splitting in such high-Z systems.

Molecular Selection Rules

Rotational Transitions

In molecular spectroscopy, pure rotational transitions occur when molecules absorb or emit radiation corresponding to changes in their rotational energy levels, typically observed in the microwave region under the rigid rotor approximation. This approximation treats the molecule as a rigid body with fixed bond lengths, allowing the rotational energy to be quantized based on the total angular momentum quantum number J. For linear molecules, such as diatomic or linear polyatomic species with a permanent electric dipole moment, the selection rule for electric dipole transitions dictates that the change in J must be \Delta J = \pm 1, prohibiting transitions where J remains unchanged or changes by more than one unit. These transitions arise from the interaction of the molecular dipole with the electric field of the radiation, leading to spectra with evenly spaced lines. The rotational constant B, defined as B = \frac{\hbar^2}{2I} where I is the moment of inertia about the rotation axis, determines the energy scale; the frequency spacing between consecutive lines is $2B, reflecting the energy difference between levels J and J+1. For symmetric top molecules, which possess two equal moments of inertia (prolate or oblate forms), an additional quantum number K describes the projection of angular momentum along the principal symmetry axis. The selection rules for electric dipole transitions depend on the orientation of the permanent dipole moment relative to the symmetry axis. For parallel transitions, where the dipole is parallel to the symmetry axis, the rules are \Delta J = \pm 1 and \Delta K = 0, ensuring that the projection along the symmetry axis remains constant. For perpendicular transitions, where the dipole is perpendicular to the symmetry axis, the rules are \Delta J = 0, \pm 1 and \Delta K = \pm 1. As a result, the rotational spectrum consists of multiple sub-bands, each corresponding to changes in K, with line spacings modulated by the rotational constants A and B (where A is associated with rotation about the symmetry axis). Line intensities in rotational spectra can exhibit alternation due to nuclear spin statistics, particularly in homonuclear diatomic molecules like . nuclei (protons) are fermions, requiring the total molecular wavefunction to be antisymmetric under nuclear exchange. This leads to two distinct forms: (total nuclear spin I=1, odd J levels) and ( I=0, even J levels), with nuclear spin degeneracies of 3 and 1, respectively. Consequently, in rotational Raman spectra of (since it lacks a for absorption), odd-J lines () are three times more intense than even-J lines (), producing an alternating 3:1 intensity pattern. At low temperatures, the para form dominates the , suppressing transitions until is reached. In contrast to microwave absorption, pure rotational Raman scattering—observed in the near-infrared or visible region—involves induced changes rather than interactions. For linear molecules, the selection rule relaxes to \Delta J = 0, \pm 2, allowing Q-branch (\Delta J = 0) lines at the frequency, as well as O-branch (\Delta J = -2) and S-branch (\Delta J = +2) lines shifted by multiples of approximately $4B. This broader rule stems from the second-order in the polarizability tensor, enabling spectra for non-polar molecules like H_2 or N_2, though intensity alternations from nuclear statistics persist in homonuclear cases. These rules align with the general selection principle of \Delta J = \pm 1 as a foundational basis, extended here by the multipolar nature of Raman processes.

Vibrational Transitions

In , vibrational transitions occur when molecules absorb or emit photons corresponding to changes in their vibrational s, primarily in the ground electronic state. For a model, the restricts transitions to those where the vibrational quantum number changes by unity, Δv = ±1, ensuring that only transitions are allowed with non-zero . This arises from the of wavefunctions, where matrix elements for Δv ≠ ±1 vanish. A vibrational mode is infrared-active only if the transition induces a change in the molecular dipole moment, such that the vibrational contribution to the dipole μ_v ≠ 0. The transition moment integral, which determines the intensity, is given by: \langle v' | \mu(q) | v \rangle \approx \left( \frac{d\mu}{dq} \right) \langle v' | q | v \rangle where q is the normal coordinate displacement, v and v' are the initial and final vibrational quantum numbers, and the derivative dμ/dq must be non-zero for the mode to be active. In practice, this means symmetric modes in centrosymmetric molecules, like the symmetric stretch in CO₂, do not change the dipole and are thus IR-inactive, while the asymmetric stretch does change it and is active. Real molecules exhibit due to deviations from the ideal parabolic potential, allowing weaker transitions with Δv = ±2, ±3, etc., though these have progressively lower intensities as the selection rule is relaxed. For polyatomic molecules with N atoms, there are 3N-6 vibrational normal modes (or 3N-5 for linear molecules), and the selection rule applies independently to each based on its and dipole derivative. In rovibrational spectra, these vibrational changes are often accompanied by rotational transitions following ΔJ = ±1.

Electronic Transitions

In molecular spectroscopy, electronic transitions involve the promotion of an electron from one to another, typically observed in the ultraviolet-visible (UV-Vis) region. The fundamental selection rules for these transitions are analogous to those in systems, requiring no change in the total (ΔS = 0) to conserve spin angular momentum during the electric dipole approximation. However, unlike isolated atoms, molecules exhibit vibronic effects where vibrational modes couple with electronic states, modifying transition intensities and sometimes relaxing strict electronic selection rules through vibronic borrowing from nearby allowed transitions. A key aspect governing the intensity of molecular electronic transitions is the Franck-Condon principle, which arises because electronic rearrangements occur much faster than nuclear motion, resulting in "vertical" transitions on the potential energy surface. The transition probability is thus determined by the overlap integral between the vibrational wavefunctions of the initial (ground) and final (excited) electronic states, with intensity proportional to |\langle \chi_{v'} | \chi_v \rangle|^2, where \chi_v and \chi_{v'} are the vibrational wavefunctions for vibrational quantum numbers v and v', respectively. This overlap is maximized when the equilibrium geometries of the ground and excited states are similar, leading to progressions in vibrational structure within electronic bands; significant geometry changes reduce the overlap for v=0 to v'=0 but favor higher vibrational levels in the excited state. In molecules, π → π* in the UV-Vis region are typically electric allowed provided the permits a change in (ungerade ↔ gerade) and adheres to ΔS = 0. For example, in conjugated systems like , the lowest-energy π → π* gains intensity only if vibronic coupling via e_{2g} modes relaxes the symmetry-forbidden nature of the pure . In coordination complexes, d-d transitions are often forbidden by the , which prohibits transitions between orbitals of the same parity in centrosymmetric environments (e.g., octahedral geometry), and by the spin selection rule if ΔS ≠ 0. These doubly forbidden transitions appear as weak bands (ε ≈ 1–100 M^{-1} cm^{-1}) due to subtle distortions or vibronic interactions that break inversion symmetry, as seen in [Ti(H_2O)_6]^{3+} where the single d-d band reflects such relaxation. Spin-forbidden examples, like those in high-spin Mn^{2+} complexes, are even weaker (ε < 1 M^{-1} cm^{-1}) unless spin-orbit coupling intervenes. Charge-transfer (CT) transitions, such as ligand-to-metal (LMCT) or metal-to-ligand (MLCT), generally obey relaxed selection rules compared to d-d transitions because they involve substantial electron density redistribution, often changing parity and leading to large geometry alterations between states. This results in intense bands (ε > 10^3 M^{-1} cm^{-1}) with broad, vibrational progressions due to poor Franck-Condon overlap at the ground-state geometry, as exemplified in permanganate ion [MnO_4^-] where LMCT occurs around 500 nm.

Advanced Applications

Multipole Transitions

In , multipole transitions describe electromagnetic interactions beyond the dominant electric (E1) approximation, arising when E1 selection rules are violated, such as due to conservation or constraints. These higher-order processes, including (M1) and electric quadrupole (E2) transitions, are significantly weaker and become relevant for "forbidden" lines in spectra where direct E1 decay is prohibited. The transition operator expands in multipole moments of the , with the order L determining the angular momentum transfer and change. Magnetic dipole (M1) transitions occur without a change in parity and allow changes in total angular momentum of \Delta J = 0, \pm 1 (excluding J=0 to J=0). These transitions are typically weaker than E1 by factors of $10^{-3} to $10^{-5}, reflecting their origin in relativistic effects like spin-orbit coupling or orbital magnetic moments. Electric quadrupole (E2) transitions conserve parity (no change) and permit \Delta J = 0, \pm 1, \pm 2 (again excluding $0 \to 0), playing a key role in fine-structure splittings where E1 is forbidden. Their rates are suppressed relative to E1 by factors involving (ka_0)^2, where k is the wave number and a_0 the Bohr radius, often on the order of $10^{-4} for optical transitions. In general, for electric multipole (EL) transitions of order L, the selection rule is |\Delta J| \leq L \leq J_i + J_f, with change given by (-1)^L; for magnetic multipole (ML), the rule is (-1)^{L+1}. These rules extend the E1 constraints (L=1, odd change) to higher orders when lower multipoles vanish. Applications include the of forbidden lines in spectra, such as the $2S \to 1S in , which proceeds via two-photon emission as a higher-order process equivalent to effective multipole contributions. Relativistic corrections, particularly the Breit interaction, incorporate transverse photon exchange between electrons, enhancing higher-multipole matrix elements and fine-structure effects in heavy atoms.

Surface Selection Rules

Surface selection rules in arise from the unique of interfaces, where the breaking of modifies probabilities for adsorbed species. In reflection-absorption (RAIRS), the surface selection rule favors vibrations with a dynamic component perpendicular to the surface, leading to enhanced signals for p-polarized light due to the enhancement to the metal surface, while s-polarized light, which couples primarily to parallel components, is strongly suppressed. This orientation sensitivity allows RAIRS to probe the of adsorbates, such as determining whether ligands are terminally bound or bridging on metal surfaces. The absence of inversion symmetry at surfaces relaxes traditional parity selection rules observed in bulk materials, activating vibrational modes in monolayers that would otherwise be infrared-inactive. For instance, in adsorbed molecular layers, the interface breaks centrosymmetry, enabling odd-parity modes with perpendicular dipole changes to become IR-active, as seen in the spectra of CO on metal surfaces where symmetric stretches gain intensity. This symmetry breaking extends molecular vibrational selection rules to two-dimensional systems at interfaces, providing insights into adsorbate-substrate interactions. In sum-frequency generation (SFG) spectroscopy, the inherent non-centrosymmetric nature of surfaces satisfies the second-order nonlinear optical requirement, with vibrational transitions obeying Δv = ±1 for resonant overlap between infrared and visible beams, thus selectively probing interfacial vibrations that are both IR- and Raman-active. High-resolution (HREELS) further illustrates surface-specific rules, where dipole scattering in specular geometry enforces a near-zero parallel transfer (Δk_parallel ≈ 0), limiting observations to long-wavelength surface phonons and adsorbate modes with minimal in-plane wavevector change. This constraint enhances sensitivity to vibrations, akin to RAIRS, but allows access to non-dipolar modes via scattering at off-specular angles. Advancements in the revealed how chiral surfaces influence -related selection rules through the chiral-induced selectivity (CISS) effect, where helical adsorbates or enantiopure modifiers on metal surfaces preferentially transmit electrons of one orientation, relaxing spin conservation in and enabling enantioselective catalysis, as demonstrated in reactions on modified surfaces.

Nuclear and Particle Physics

In , selection rules govern transitions between nuclear states via gamma emission, where the change in nuclear spin satisfies ΔI = 0, ±1 for the dominant radiation, with the no 0 → 0 transition rule prohibiting direct emission between spin-zero states./07%3A_Radioactive_Decay_Part_II/7.01%3A_Gamma_Decay) For electric (E1) transitions, the selection rule requires a change in , Δπ = -1, ensuring the emitted photon's multipolarity aligns with conservation laws./07%3A_Radioactive_Decay_Part_II/7.01%3A_Gamma_Decay) These rules, rooted in and conservation, parallel atomic selection rules (ΔJ = 0, ±1) but apply to collective nuclear excitations at MeV scales./07%3A_Radioactive_Decay_Part_II/7.01%3A_Gamma_Decay) Isospin conservation plays a key role in nuclear interactions, with the strong and electromagnetic forces preserving total isospin (ΔT = 0), while the weak force allows ΔT = 0 or 1, as seen in beta decay transitions where Fermi operators enable isoscalar changes and Gamow-Teller operators isovector shifts. In particle decays, such as beta decay, angular momentum conservation imposes ΔJ = 0, ±1 for allowed transitions, explicitly forbidding 0 → 0 due to the lepton pair's inability to carry zero net angular momentum without orbital contributions. Parity is maximally violated in these weak processes, as described by the vector-axial vector (V-A) theory proposed in 1957, which unifies the interaction form across fermions and explains the observed asymmetry in decay distributions. Notable examples illustrate these rules in practice. The enables recoil-free gamma emission in solids for transitions with ΔI = 0, such as certain (M1) decays, allowing measurements without thermal broadening. In neutron capture reactions, selection rules on and determine the formation of compound nuclear states, with s-wave captures favoring even parity and influencing cross-sections for astrophysical processes like the . In modern , selection rules for hadronic decays are formulated within quark models, predicting allowed channels based on symmetry and , such as forbidden transitions in charmonium states. These models have been refined post-2010 through simulations, which compute decay matrix elements and validate selection rule violations in exotic hadrons, providing quantitative insights into dynamics.

References

  1. [1]
    Selection Rules | Modeling and Experimental Tools with Prof. Magnes
    Jan 4, 2019 · The selection rules describe whether a quantum transition is considered to be forbidden or allowed. This means that not all transitions of electrons are ...Missing: definition | Show results with:definition
  2. [2]
    A Primer on Quantum Numbers and Spectroscopic Notation
    Contents · Quantum Numbers · Quantum Numbers for Atoms · Spectroscopic Notation · Terms, Configurations, and Levels · Selection Rules · Multiplets. M Degeneracy and ...
  3. [3]
    [PDF] UNDERSTANDING SELECTION RULES IN ATOMIC ...
    Mar 27, 2014 · Most modern expositions of applied group theory in quantum mechanics derive selection rules as an illustrative example of group theoretic ...
  4. [4]
    The quantum theory of the emission and absorption of radiation
    Dirac Paul Adrien Maurice. 1927The quantum theory of the emission and absorption of radiationProc. R. Soc. Lond. A114243–265http://doi.org/10.1098/rspa ...
  5. [5]
    [PDF] Quantum Physics III Chapter 4: Time Dependent Perturbation Theory
    Apr 4, 2022 · We will use the known Hamiltonian H(0) to define some Heisenberg operators and the perturbation δH will be used to write a Schrödinger equation.
  6. [6]
    [PDF] DAMTP - 9 Perturbation Theory II: Time Dependent Case
    are obeyed. These selection rules were derived under the assumption that the transition is just due to the dipole approximation where the perturbation involves ...
  7. [7]
    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 ...
  8. [8]
    [PDF] Time-Dependent Perturbation Theory and Molecular Spectroscopy
    Thus, the vibrational selection rule is ∆v = 0, ±1. This selection rule is rigorous pro- vided that higher-order terms in the expansion of the dipole-moment ...
  9. [9]
    [PDF] Chapter 6 Groups and Representations in Quantum Mechanics
    The application of group theory to selection rules necessitates the introduction of the “direct product” of matrices and groups, though here, too, quantum ...
  10. [10]
    Spectroscopic Selection Rules: The Role of Photon States
    Selection rules play a central role in spectroscopy. In an introductory undergraduate course it is frequently desirable to give some justification for these ...
  11. [11]
    Atomic Spectros. - Spectral Lines - NIST
    Selection rules for discrete transitions. Electric dipole (E1) ("allowed"), Magnetic dipole (M1) ("forbidden"), Electric quadrupole (E2) ("forbidden") ...
  12. [12]
    Lasers
    ∆l = ±1 is called a selection rule. If this selection rule is not satisfied, an optical transition is forbidden (unlikely). The selection rule ∆l = ±1 is ...<|control11|><|separator|>
  13. [13]
    [PDF] Atomic Spectra in Astrophysics
    Electric dipole selection rules are two types: rigorous rules - must always be obeyed; propensity rules - lead to weaker transitions. Rigorous selection rules.
  14. [14]
    Molecular Spectroscopy - an overview | ScienceDirect Topics
    Selection rules govern what transitions can occur (are allowed) and what transitions cannot occur (are forbidden). It is also based on the symmetries of the ...<|separator|>
  15. [15]
    [PDF] 4. Atoms in Electromagnetic Fields - DAMTP
    More powerful selection rules come from looking at the other angular momentum quan- tum numbers. Neglecting spin, the atomic states | i are labelled by |n ...
  16. [16]
    Coherence and resonance effects in the ultra-intense laser-induced ...
    Jan 6, 2016 · Discrepancies between master equation and rate equation are the results of coherence-induced Rabi oscillations and power-broadening effects.<|control11|><|separator|>
  17. [17]
    Electric Dipole Approximation and Selection Rules
    There is one absolute selection rule coming from angular momentum conservation, since the photon is spin 1. No $j=0$ to $j=0$ transitions in any order of ...Missing: ΔJ = | Show results with:ΔJ =
  18. [18]
    [PDF] The Wigner-Eckart Theorem - University of California, Berkeley
    Thus, the selection rule m = m′ + q means that q is the z-component of the spin of the emitted photon, so that the z-component of angular momentum is conserved ...
  19. [19]
    Selection rules for electric dipole transitions - Oxford Academic
    These selection rules allow one to calculate quickly which transitions will occur once the angular momentum quantum numbers of the energy levels are specified.
  20. [20]
    [PDF] Lecture 19, Hydrogen Atom - DSpace@MIT
    Thus, if the photons are z- polarized, they don't change the z-projection of the system angular momentum, so Δm=0. Meanwhile x- or y-polarized photons have ±1 ...
  21. [21]
    [PDF] Selection rule for electric dipole transition - bingweb
    Dec 2, 2024 · The angular momentum of the photon can be derived from the conservation of the total angular momentum.Missing: ΔJ | Show results with:ΔJ
  22. [22]
    Polarization-dependent intensity ratios in double resonance ...
    Nov 8, 2023 · ΔM = ±1 selection rule). We treat the pump transition as a separate two-level system for each M value. The steady-state change in population ...<|control11|><|separator|>
  23. [23]
    [PDF] Lecture 21: The Parity Operator
    The parity operator is defined, is Hermitian, and its own inverse. It is related to parity inversion, where f(-x) = f(x) defines symmetry.
  24. [24]
    [PDF] 7.1 Dipole Transitions
    The ladder operators change the z-component of their corresponding angular momentum by 1, and therefore we know that ∆l = 0, ∆ml = ±1, and ∆ms = ±1. The ...
  25. [25]
    7.24: Selection rules - Physics LibreTexts
    Mar 5, 2022 · In all of these situations there are selection rules that restrict certain quantum numbers from changing by other than certain amounts.
  26. [26]
    [PDF] parity.pdf
    As noted above, Laporte's rule is equivalent to the statement that the intrinsic parity of the photon is negative. In the case of the matrix element (51), the ...
  27. [27]
    [PDF] Experimental Evidence for Particle Intrinsic Parity
    This is actually the reason that the photon has negative parity, and is referred to as an 1− state. Photons are emitted via atomic dipole transitions where.
  28. [28]
    7.4 Conservation Laws in Emission
    In such transitions the photon comes out with one unit of angular momentum, i.e. must change from odd to even or vice-versa. It cannot stay the same.Missing: ΔJ | Show results with:ΔJ
  29. [29]
    Experimental Test of Parity Conservation in Beta Decay | Phys. Rev.
    Experimental test of parity conservation in beta decay. CS Wu, E. Ambler, RW Hayward, DD Hoppes, and RP Hudson. Columbia University, New York, New York.
  30. [30]
    [PDF] line radiation and spectroscopy
    Radiative Transitions. The strongest lines correspond to electric dipole transitions (Rybicki and Lightman. 1979, pp. 271ff). These are the lowest order terms ...<|control11|><|separator|>
  31. [31]
    [PDF] APAS 5110. Internal Processes in Gases. Fall 1999. - JILA
    The selection rules for magnetic dipole transitions follow from equation (3.5): only transi- tions between states in which the matrix elements of either L or S ...
  32. [32]
    [PDF] Spectroscopy 2: electronic transitions The energies needed to ...
    - transitions are allowed. 2) The Laporte selection rule: The only allowed transitions are transitions that are accompanied by a change in parity. u → g ...
  33. [33]
    [PDF] Astro 501: Radiative Processes Lecture 31
    Apr 8, 2013 · Laporte's rule: no transitions between two states of the same parity. Q: what is a parity transformation? Q: why is ~d fi. = 0 if i and j have ...<|control11|><|separator|>
  34. [34]
    [PDF] Atomic Physics - Doyle Group | Harvard
    atoms in the -LS-coupling scheme usually leads to an interval rule. Fìrrther examples of energy levels are given in the exercises at the end of this chapter ...<|control11|><|separator|>
  35. [35]
    Atomic Spectroscopy - A Compendium of Basic Ideas, Notation ...
    Jul 9, 2009 · Atomic Spectroscopy - A Compendium of Basic Ideas, Notation, Data, and Formulas ...
  36. [36]
    Hund's Rules for Atomic Energy Levels - HyperPhysics
    Hund's Rule #1: The term with the maximum multiplicity lies lowest in energy. The explanation of the rule lies in the effects of the spin-spin interaction.
  37. [37]
  38. [38]
    13.12: The Selection Rule for the Rigid Rotor
    ### Summary of Selection Rules for Rotational Transitions in Rigid Rotor Molecules
  39. [39]
    Rotational Spectra of Rigid Rotor Molecules - HyperPhysics
    For a rigid rotor diatomic molecule, the selection rules for rotational transitions are ΔJ = +/-1, ΔMJ = 0 . The rotational spectrum of a diatomic molecule ...
  40. [40]
    [PDF] 5.80 Small-Molecule Spectroscopy and Dynamics
    For a symmetric top, the selection rules for J and M are identical to those for a linear molecule, namely: ∆J = ± 1 ∆M = 0 for radiation polarized along the z ...
  41. [41]
    [PDF] ( ) ( )e
    The selection rules for a rotational transition in this case are ΔJ=±1 and ΔK=0, and thus eqs. 13-11 and 13-12 continue to be valid for absorption frequencies ...
  42. [42]
    [PDF] Spectroscopy 1: rotational and vibrational spectra
    We have treated molecules as rigid rotors. However ... Rotational Raman spectra. The gross selection rule: the molecule must be anisotropically polarizable.
  43. [43]
    [PDF] Rotational Raman Spectra of Diatomic Molecules
    Nov 1, 2010 · j+2. - E j. = (ħ2/2I)(4j + 6), for Stokes Raman scattering (Δj = +2), or. ΔE = E j. - E j-2. = (ħ2/2I)(4j - 2), for anti‐Stokes Raman scattering ...
  44. [44]
    13.5: Vibrational Overtones - Chemistry LibreTexts
    Oct 16, 2025 · Anharmonic Oscillator Selection Rules​​ Overtones occur when a vibrational mode is excited from to (the first overtone) or to (the second ...
  45. [45]
    Vibrational Transition Energies For An Anharmonic Diatomic Molecule
    In infrared spectroscopy, transitions are allowed between adjacent v levels for a harmonic oscillator - the selection rule is: where an increase in v ...
  46. [46]
    15.2: Vibration-Rotation Transitions - Chemistry LibreTexts
    Apr 28, 2023 · The Dipole Moment Derivatives​​ This result can be interpreted as follows: Each independent vibrational mode of the molecule contributes to the μ ...The Dipole Moment Derivatives · Rotational Selection Rules
  47. [47]
    [PDF] Vibrational Transition Moments and Dipole Derivatives References
    Nov 9, 2006 · We may compute the electric-dipole vibrational transition moment beginning from the Born-Oppenheimer approximation, in which we assume that ...
  48. [48]
    Vibrational Modes of Carbon Dioxide
    C-O asymmetric stretching 2565 cm-1 (IR intensity = 1.0) (Raman inactive) ... C-O symmetric stretching 1480 cm-1 (IR inactive) (Raman active). Jmol ...
  49. [49]
    Infrared Spectroscopy - Chemistry LibreTexts
    Jan 29, 2023 · It has 4 modes of vibration (3(3)-5). CO2 has 2 stretching modes, symmetric and asymmetric. The CO2 symmetric stretch is not IR active ...
  50. [50]
    Infrared Absorption – Short Stories in Instrumental Analytical ...
    For a true quantum mechanical harmonic oscillator, there is a selection rule that vibrational transitions must be between adjacent vibrational levels. The ...12 Infrared Absorption · Spectral Region · Ir-Active Vibrations
  51. [51]
    14: Harmonic Oscillators and IR Spectroscopy - Chemistry LibreTexts
    Dec 5, 2019 · We introduced a qualitative discussion of IR spectroscopy and then focused on selection rules for what vibrations are IR-active and can be seen in IR spectra.
  52. [52]
    [PDF] CHM 305 - Lecture 11 - Molecular Spectroscopy
    These rules that determine which states can be reached by optical transitions are called selection rules. An additional rule for infrared spectroscopy is that ...
  53. [53]
    [PDF] Selection rules and transition moment integral
    Selection rules have been divided into the electronic selection rules, vibrational selection rules (including. Franck-Condon principle and vibronic coupling), ...
  54. [54]
    [PDF] Lecture 30 and part 31: Electronic Spectroscopy. Franck-Condon ...
    The Franck-Condon principle is based on sudden promotion of one e- , so fast that nuclei respond only after the e- excitation. ΔR = 0 “vertical transitions”. ΔP ...
  55. [55]
    Franck-Condon Principle - an overview | ScienceDirect Topics
    The Franck–Condon principle refers to the concept that during electronic transitions caused by photon absorption, the positions and velocities of nuclei ...
  56. [56]
    [PDF] UV/VIS Spectroscopy
    Of course in this latter case no n→π∗ transitions are allowed. Only π→π∗ transitions are possible since the molecule contains no lone pair (n). Page 4 ...
  57. [57]
    Photochemistry: absorbance rules - csbsju
    Decide whether each of the following transitions would be allowed by symmetry. a) π → π* b) p → π* c) p → σ* d) d → d. Symmetry selection rules are in ...
  58. [58]
    [PDF] Color of Transition Metal Complexes
    Laporte forbidden but also spin forbidden. Absorptions that are doubly forbidden transitions are extremely weak. It is understandable, then, that dilute ...
  59. [59]
    Magnetic Dipole Transitions - Richard Fitzpatrick
    Jan 22, 2016 · Hence, the selection rules for magnetic dipole transitions are. $\displaystyle {\mit\Delta} l$, $\displaystyle =0,$, (8.205). $\displaystyle ...Missing: ΔJ | Show results with:ΔJ
  60. [60]
    A.25 Multipole transitions
    This condition is called the parity “selection rule” for an electric dipole transition. If it is not satisfied, the electric dipole contribution is zero and a ...
  61. [61]
    Electric Quadrupole Transitions - Richard Fitzpatrick
    According to Equation (8.195), the quantity that mediates spontaneous electric quadrupole transitions is $\displaystyle \mbox{\boldmath$\epsilon$}$Missing: ΔJ | Show results with:ΔJ
  62. [62]
    [PDF] ALL-ORDER METHODS FOR RELATIVISTIC ATOMIC STRUCTURE ...
    The row labeled B(2) is the contribution from the second-order Breit interaction. For atoms with two electrons beyond a closed core, it is of considerable inter ...
  63. [63]
    Breit corrections to individual atomic and molecular orbital energies
    Jan 29, 2018 · The Breit interaction gives the leading relativistic correction to the instantaneous Coulomb electron–electron interaction potential in quantum ...Missing: multipoles | Show results with:multipoles
  64. [64]
    Symmetry and the Surface Infrared Selection Rule for the ...
    Symmetry and the surface infrared selection rule for the determination of the structure of molecules on metal surfaces.Missing: vibrations | Show results with:vibrations
  65. [65]
    Tutorials in vibrational sum frequency generation spectroscopy. I ...
    Jan 20, 2022 · This tutorial series is aimed at people entering the VSFG world without a rigorous formal background in optical physics or nonlinear spectroscopy.Missing: Δv=± | Show results with:Δv=±
  66. [66]
    Chiral Induced Spin Selectivity Gives a New Twist on Spin-Control in ...
    Oct 12, 2020 · A new phenomenon was discovered that relates the electron's spin to the handedness of chiral molecules and is now known as the chiral induced spin selectivity ...
  67. [67]
    [PDF] 15. Nuclear Decay - Particle and Nuclear Physics Prof. Tina Potter
    In general, EL transitions Parity = (−1). L. ML transitions Parity = (−1). L+ ... Selection rules; Fermi, Gamow-Teller; allowed, forbidden. γ-decay.
  68. [68]
    Investigation of the Δn = 0 selection rule in Gamow-Teller transitions
    Jun 10, 2019 · The selection rules for allowed β decay are total angular momentum change Δ I = 0 , ±1 and no parity change between the initial (decaying) and ...
  69. [69]
    [PDF] V-A: Universal Theory of Weak Interaction
    He suggested that the weak decay interaction was of the universal form V-A with maximal parity violation, in which every field was multiplied by the chiral.
  70. [70]
    Hyperfine interaction studies in the David Shirley group, 1960–1975 ...
    Jun 8, 2022 · ... ΔI = 0 decays (no change in nuclear spin and no change in parity) ... Perturbed angular correlations and Mössbauer effect. J. Vac. Sci ...<|control11|><|separator|>
  71. [71]
    Test of the SO(6) selection rule in 196Pt using cold-neutron capture
    In order to test this selection rule, absolute B(E2) values between σ = N − 2 and states need to be measured. This is not an easy task as the first σ = N − 2 ...
  72. [72]
    Selection Rules for Hadronic Transitions of Mesons | Phys. Rev. Lett.
    Jun 3, 2014 · Selection rules for hadronic transitions are derived and used to identify mesons that are candidates for ground-state energy levels in the BO potentials.