Fact-checked by Grok 2 weeks ago

Hund's rules

Hund's rules are a set of three empirical guidelines in , formulated by German physicist in 1925, that predict the relative energies of electronic states in multi-electron atoms by determining the ground-state based on the total spin, orbital, and angular momenta of the electrons. These rules apply primarily under the Russell-Saunders (L-S) coupling scheme, where the orbital angular momenta of individual electrons couple to form a total orbital angular momentum L, the spin angular momenta couple to a total spin S, and these then couple to a total angular momentum J. They provide a systematic way to order atomic energy levels from spectroscopic data and are fundamental for understanding atomic spectra, electronic configurations, and chemical bonding in transition metals and rare-earth elements. The three rules are stated as follows: Physically, Hund's rules stem from the and the minimization of electron repulsion in degenerate orbitals. The first rule is explained by the antisymmetric spatial wavefunction for parallel spins, which keeps electrons farther apart on average, lowering the energy compared to symmetric spatial wavefunctions for paired-spin states. The second rule reflects the tendency for electrons to occupy orbitals with higher azimuthal quantum numbers when possible, increasing their average radial distance. The third rule accounts for the spin-orbit coupling energy, which is negative when the spin and orbital moments are antiparallel for light atoms. These principles were derived empirically from early spectroscopic observations but have been justified quantum mechanically through Hartree-Fock approximations and configuration interaction methods. Hund's rules are essential in fields like and , enabling the prediction of magnetic properties, such as in atoms with unpaired s, and the stability of molecular orbitals in compounds. For example, in the carbon atom's (1s² 2s² 2p² ), the rules select the ³P as lowest, with two unpaired p-s contributing to its triplet multiplicity. Exceptions occur in heavy atoms where relativistic effects and j-j dominate, or in ions with strong electron correlation, but the rules hold accurately for most light elements and transition metals up to the fourth row. Ongoing research uses to explore violations and generalizations, such as in quantum dots or molecular systems.

Background and Context

Historical Development

The development of Hund's rules emerged during a pivotal period in , amid efforts to interpret complex atomic spectra using the framework of the . In 1924, British physicist Edmund C. Stoner published his seminal paper "The Distribution of Electrons among Atomic Levels" in the , where he proposed a systematic arrangement of electrons into subshells based on spectroscopic observations and magnetic properties, effectively anticipating aspects of the by incorporating a fourth related to . Stoner's work provided an empirical foundation for understanding electron configurations in atoms, influencing subsequent theoretical advancements in . Building directly on Stoner's insights and the growing body of data, German physicist formulated the core principles of what would become known as Hund's rules in 1925. In his "Zur Deutung verwickelter Spektren, insbesondere der Elemente bis ," published in Zeitschrift für Physik, Hund introduced guidelines for determining the lowest-energy term symbols of atomic configurations by prioritizing maximum multiplicity and orbital , derived from analyses of spectra. A follow-up later that year in the same journal refined these ideas, extending them to broader periodic trends observed in line spectra. These proposals were empirically grounded in the observed regularities of atomic spectra, which revealed patterns in splittings that classical models could not explain. Hund's contributions from 1925 to 1927 coincided with the rapid evolution from the —characterized by Bohr-Sommerfeld quantization rules—to the matrix mechanics of in 1925, the wave mechanics of in 1926, and Paul Dirac's relativistic quantum formulation in 1928. Working in under , Hund integrated these emerging quantum concepts into his analyses of atomic and molecular systems, applying them to development, particularly in his 1927 book Linienspektren und periodisches System der Elemente, which synthesized spectral data with quantum postulates to explain periodic properties and bonding. This work bridged the gap between semi-classical models and full , establishing Hund's rules as a cornerstone for predicting ground-state multiplicities in multi-electron atoms during this transformative era.

Physical Basis

The Pauli exclusion principle dictates that no two electrons in an atom can occupy the same , thereby restricting possible configurations for electrons in degenerate orbitals and ensuring an antisymmetric total wavefunction for the multi-electron system. This fermionic requirement forces electrons with parallel to reside in different spatial orbitals, which spatially separates them and reduces the average electron-electron distance compared to configurations with paired . The exchange interaction, arising from the antisymmetrization of the wavefunction required by the Pauli principle, plays a central role in favoring configurations with maximum total spin multiplicity. For electrons with parallel spins, the exchange term in the two-electron Coulomb interaction contributes a negative energy shift, effectively lowering the repulsion energy relative to antiparallel spin configurations where such exchange is absent. The total spin angular momentum for equivalent electrons is given by \mathbf{S} = \sum_i \mathbf{s}_i, where \mathbf{s}_i is the spin operator for the i-th electron (with |\mathbf{s}_i| = \frac{1}{2} \hbar). Maximizing the orbital similarly lowers the energy by allowing electrons to avoid each other more effectively through increased angular nodal structure in the wavefunctions. States with higher L exhibit greater azimuthal variation, which minimizes the repulsion by keeping electron densities apart on average, an effect sometimes termed orbit-orbit . The orbital is \mathbf{L} = \sum_i \mathbf{l}_i, where \mathbf{l}_i is the orbital for the i-th electron. Spin-orbit coupling acts as a smaller perturbative effect that couples individual spins and orbits, leading to splitting in energy levels. This relativistic is described by the \hat{H}_{SO} = \sum_i \zeta(r_i) \mathbf{l}_i \cdot \mathbf{s}_i, where \zeta(r_i) is the coupling strength depending on the radial coordinate, and it influences the ordering of states with different total J = L + S but is typically weaker than exchange or electrostatic effects in light atoms.

Formulation of the Rules

Rule 1: Maximum Spin Multiplicity

Hund's first rule states that, for a given of an , the state characterized by the maximum total S possesses the lowest energy among the possible Russell-Saunders terms. This prioritization ensures that the ground state maximizes the number of unpaired electrons with parallel spins, as observed empirically in atomic spectra during the development of in the . The physical rationale for this rule lies in the reduction of electron-electron repulsion when spins are parallel. According to the , the total wavefunction of fermions must be antisymmetric under particle exchange. For parallel spins (symmetric spin part), the spatial wavefunction must be antisymmetric, which spatially separates the electrons and minimizes their repulsion. In contrast, antiparallel spins (antisymmetric spin part) require a symmetric spatial wavefunction, leading to greater overlap and higher repulsion energy. This effect is quantified in the two-electron case through the two-electron , where the interaction term separates into a direct integral C = \langle 12 | 1/r_{12} | 12 \rangle and an exchange integral K = \langle 12 | 1/r_{12} | 21 \rangle, with K > 0. The (S = 1, parallel spins) has energy C - K, while the (S = 0, antiparallel spins) has C + K, making the high-spin state lower in energy by $2K. The spin multiplicity, denoted as $2S + 1, provides a convenient label for the degeneracy of the spin states; for instance, S = 1 yields a multiplicity of 3 (triplet), and S = 0 yields 1 (). This notation highlights how the rule favors states with higher multiplicity for the . To illustrate, consider the p^2 , as in the neutral carbon . The possible terms are ^3P (S = 1), ^1D (S = 0), and ^1S (S = 0); according to the first rule, the ^3P term lies lowest in energy, forming the of carbon.

Rule 2: Maximum Orbital Angular Momentum

The second Hund's rule specifies that, among the possible Russell-Saunders terms arising from a given that possess the maximum total S, the term with the largest total orbital quantum number L exhibits the lowest . This preference for maximum L stems from the nature of electron-electron interactions in atoms. When the individual orbital angular momenta \vec{l}_i of the electrons are aligned to yield a high total L, the electrons tend to orbit the in a correlated manner that maximizes their average spatial separation. Consequently, the repulsive interactions between electrons are minimized, leading to a lower overall compared to states with smaller L, where electrons are more likely to occupy overlapping regions of space. The total orbital \vec{L} results from the vector sum of the individual orbital angular momenta: \vec{L} = \sum_i \vec{l}_i. The magnitude L can take integer values from \left| \sum l_i \right| down to 0 or 1 (depending on ), but the maximum L corresponds to the "stretched" configuration where all \vec{l}_i are aligned as much as possible, subject to the . In , terms are labeled ^{2S+1}L, where L is represented by letters: S for L=0, P for L=1, D for L=2, F for L=3, G for L=4, and so forth. For instance, in the d^2 electron configuration of equivalent electrons (as in the titanium atom), the possible terms with maximum S=1 (triplet multiplicity) are ^3F (L=3) and ^3P (L=1). The second rule predicts that the ^3F term has lower energy than the ^3P term, which is confirmed experimentally as the ground state.

Rule 3: Spin-Orbit Interaction

The third Hund's rule specifies the ordering of fine-structure levels within a spectroscopic term arising from the coupling of total orbital angular momentum \vec{L} and total spin angular momentum \vec{S} to form the total angular momentum \vec{J}. For an unfilled subshell containing fewer electrons than needed to half-fill it (i.e., n < 2\ell + 1), the level with the minimum J has the lowest energy. For a subshell more than half full (n > 2\ell + 1), the level with the maximum J has the lowest energy. For a exactly half-filled subshell (n = 2\ell + 1), the ground level has J = L + S. This rule stems from the spin-orbit interaction, which splits the degenerate ^{2S+1}L term into $2J+1 levels for each possible J ranging from |L - S| to L + S in integer steps. The interaction energy is given by the expectation value \langle H_{SO} \rangle = A \vec{L} \cdot \vec{S}, where A is the spin-orbit coupling constant. Using the identity \vec{L} \cdot \vec{S} = \frac{1}{2} [J(J+1) - L(L+1) - S(S+1)], the energy shift becomes \frac{A}{2} [J(J+1) - L(L+1) - S(S+1)]. For subshells less than half full, A > 0 (normal coupling), so the minimum J yields the lowest energy; for more than half full, A < 0 (inverted coupling), favoring maximum J. This behavior aligns with the Landé interval rule, where the separation between consecutive J levels is proportional to J+1, with the sign determining the ordering. A representative example is the ground state of the oxygen atom, with configuration $1s^2 2s^2 2p^4. The $2p^4 subshell yields the ^3P term (L=1, S=1), so possible J values are 0, 1, and 2. Since n=4 > 3 (half-filling for p subshell), the inverted coupling applies, and the ^3P_2 level is the lowest energy, as confirmed by spectroscopic data.

Applications and Examples

Ground State Configurations

To determine the ground state term symbol for an atomic configuration using Hund's rules, first identify all possible terms ^{2S+1}L arising from the equivalent electrons in the incomplete subshell by enumerating allowed microstates that satisfy the . Then, apply Rule 1 to select the term with the maximum spin multiplicity $2S+1, where S is the total . Among terms with the same multiplicity, apply Rule 2 to choose the one with the maximum orbital L. Finally, apply Rule 3 to determine the total J for the ground level: for subshells less than half full, select the minimum J = |L - S|; for subshells more than half full, select the maximum J = L + S; for exactly half full, J = S. For the carbon atom with configuration $1s^2 2s^2 2p^2, the possible terms from the p^2 subshell are ^1S, ^1D, and ^3P. Rule 1 selects the triplet term ^3P (S=1) over the singlets. Rule 2 confirms ^3P (L=1) as the highest L for multiplicity 3. Since p^2 is less than half full (half is p^3), Rule 3 gives J = |1 - 1| = 0, yielding the ground state ^3P_0. For the nitrogen atom with configuration $1s^2 2s^2 2p^3, the possible terms from p^3 are ^2D, ^2P, and ^4S. Rule 1 selects ^4S (S=3/2). With L=0, Rules 2 and 3 are satisfied with J = 3/2, giving ^4S_{3/2}. For neutral iron with configuration [\mathrm{Ar}] 3d^6 4s^2, the closed $4s^2 subshell contributes no , so the terms arise from the d^6 electrons. The possible terms include ^5D, ^3H, ^3F, among others. Rule 1 selects the quintet terms (S=2), and Rule 2 chooses ^5D (L=2) as the highest L. Since d^6 is more than half full (half is d^5), Rule 3 gives J = 2 + 2 = 4, yielding ^5D_4. The ground state terms for equivalent p^n (n=1–6) and d^n (n=1–10) configurations, determined by Hund's rules (without J, as J depends on the specific subshell filling), are summarized in the following table. Note that configurations with n > 2l+1 electrons are equivalent to holes with (2l+1 - n) electrons, yielding the same terms but potentially inverted J ordering.
Subshelln=1 (or 9/5)n=2 (or 8/4)n=3 (or 7)n=4 (or 6)n=5n=6n=10
p^n^2P^3P^4S^3P^2P^1S
d^n^2D^3F^4F^5D^6S^5D^1S
These predicted ground states are verified by atomic spectra, where the lowest-energy levels correspond to the Hund's rule terms. For instance, the ^3P_0 level of carbon lies at 0 cm⁻¹, with excited components ^3P_1 at 16.4 cm⁻¹ and ^3P_2 at 43.5 cm⁻¹, and higher terms like ^1D_2 at 10,176 cm⁻¹. Similarly, nitrogen's ^4S_{3/2} is at 0 cm⁻¹, with the next term ^2D_{5/2} at 19,200 cm⁻¹. For iron, the ^5D_4 level is the ground state at 0 cm⁻¹, consistent with observed transitions in the visible spectrum.

Excited State Configurations

Hund's rules apply to configurations in atoms when the terms arise from a single electronic configuration with degenerate orbitals, allowing the rules to order the relative energies of the LS terms and their components within that configuration. For such cases, the first rule prioritizes states with maximum spin multiplicity (2S+1), the second favors maximum orbital L, and the third determines the ordering of total J levels based on whether the subshell is less than or half-filled (lowest J lowest energy) or more than half-filled (highest J lowest energy). However, in s, the rules' predictions can be less reliable than for ground states due to interconfiguration interactions and other perturbations that mix terms from nearby configurations, potentially inverting expected orderings. A classic example is the 1s2p excited configuration of helium, where the triplet ^3P term (S=1, L=1) lies below the singlet ^1P term (S=0, L=1), consistent with the first rule favoring higher multiplicity. The ^3P term further splits into fine structure levels ^3P_2 at 169086.77 cm^{-1}, ^3P_1 at 169086.84 cm^{-1}, and ^3P_0 at 169087.83 cm^{-1}, while the ^1P_1 level is at 171134.60 cm^{-1}; for this non-equivalent electron configuration, the ordering is inverted relative to Hund's third rule prediction for less-than-half-filled subshells (which would place lowest J lowest), with J=2 lowest in energy. Another illustration is the 3p ^2P excited state of sodium, where the configuration (less than half-filled) yields ^2P_{1/2} at 16956 cm^{-1} and ^2P_{3/2} at 16973 cm^{-1}, adhering to the third rule with J=1/2 lower than J=3/2. Unlike ground states, where Hund's rules often suffice to identify the absolute lowest energy term across configurations, excited states require more complete calculations—such as those incorporating electron correlation or configuration interaction—to determine precise excitation energies and overall spectral positions, as the rules only govern intra-configuration ordering. These predictions nonetheless play a crucial role in interpreting atomic spectra by identifying the dominant terms responsible for observed lines, enabling the assignment of allowed electric dipole transitions (ΔS=0, ΔL=±1, ΔJ=0,±1 with parity change) and the fine structure splittings that resolve multiplets in emission or absorption spectra. For instance, in helium's spectrum, the rules explain why triplet transitions like 1s2p ^3P to 1s2s ^3S are prominent in certain astrophysical contexts.

Limitations and Extensions

Exceptions in Light Atoms

In light atoms, where relativistic effects are minimal, Hund's rules generally provide accurate predictions for ground-state term symbols due to the dominance of LS coupling and weak spin-orbit interactions. However, exceptions arise primarily in excited states, where strong configuration interaction () between nearly degenerate configurations can perturb the energy ordering, leading to violations of the first or second rules. These deviations highlight the limitations of the single-configuration approximation underlying Hund's rules, as electron correlation allows mixing that stabilizes lower-multiplicity or lower-L terms over their expected counterparts. A prominent example occurs in the neutral beryllium atom (Be I), for the excited 2p² configuration. According to Hund's first rule, the term with maximum multiplicity, ^3P (2S+1=3), should have the lowest , followed by the singlets ^1D and ^1S. Experimental reveals, however, that the ^1D term lies below the ^3P term, with the ^1D level at approximately 56882 cm⁻¹ and the ^3P at 59694 cm⁻¹ relative to the . This inversion violates the maximum multiplicity rule. The cause is intense between the 2p² ^1D configuration and the nearby 2s ^1D Rydberg series (particularly n=3 and higher), which lowers the of the singlet while repelling the triplet upward. Such mixing is facilitated by the similar radial distributions of the 2s and 2p orbitals in light atoms like , enhancing correlation effects. Similar, though less pronounced, perturbations appear in excited states of other atoms, such as (B I). For the 2s 2p² configuration, the high-multiplicity ^4P (from parallel spins of the three electrons) correctly lies lowest at around 28870 cm⁻¹ above the ground ^2P° state, consistent with Hund's first rule. However, within the subshells, with configurations like 2s² np perturbs the singlet-triplet separations in associated , occasionally inverting expected orderings for higher excited levels as observed in vacuum ultraviolet spectra. In (N I), the ground p³ ^4S adheres to the rules, but excited p² ns or p² np configurations show CI-induced shifts in L ordering, where lower-L drop below higher-L ones due to interactions with continuum states. These cases underscore how electron correlation, rather than , drives exceptions in light atoms (Z ≤ 10). Experimental confirmation comes from high-resolution , including and laser-based techniques, which resolve fine details of these perturbed levels. For instance, UV absorption and spectra of Be I reveal the anomalous ^1D placement through transitions from the 2s 2p ^3P state, with wavelengths matching the inverted energies to within 0.1 cm⁻¹. In alkali atoms like sodium (Na I), excited states involving multi-electron configurations (e.g., 3p nd or 4s 4p) exhibit analogous CI effects; the expected triplet terms are sometimes destabilized by mixing with singlets from nearby Rydberg manifolds, leading to irregular multiplet splittings observed in selective excitation spectra. These breakdowns, while not altering ground states, emphasize the need for multi-configuration treatments in precise atomic modeling.

Relation to Modern Quantum Chemistry

In modern quantum chemistry, Hund's rules serve as qualitative guides for constructing configuration interaction (CI) expansions in post-Hartree-Fock methods, particularly spin-adapted CI, where they enforce high-spin configurations to introduce sparsity and reduce the number of significant configuration state functions. For instance, in systems like the [CaMn₃(IV)O₄] , CAS(9,9) expansions correlate nine electrons in t2g orbitals, and applying Hund's first rule enhances sparsity by prioritizing parallel-spin alignments. In the [Fe₄(III)S₄] , this approach decreases the number of significant CI coefficients from 170 to 50, improving computational efficiency while prioritizing energetically favorable parallel-spin alignments. This integration validates the rules' role in selecting dominant terms for accurate energy calculations beyond mean-field approximations. Density functional theory (DFT) incorporates Hund's principles through corrections that mitigate self-interaction errors (SIE), which otherwise violate the rules by delocalizing electrons and underestimating and orbital polarization. In DFT+U methods, a Hubbard parameter (U ≈ 8–10 eV) on localized orbitals, combined with manual enforcement of 4f occupancy matrices, restores compliance with Hund's second rule, enabling correct ground-state predictions in rare-earth compounds like TbMn₆Sn₆, where non-enforced calculations raise the energy by 340 meV. Self-interaction corrections (SIC) and orbital polarization corrections (OPC) also aim to maximize orbital moments per Hund's guidelines, though DFT+U proves most effective for (MA) in systems such as RCo₅ and R₂Fe₁₄B. Meta-GGA functionals, by including density, indirectly support higher- states in open-shell atoms without explicit enforcement, aligning with benchmarks where Hund's rules hold for single-atom configurations. Extensions of Hund's rules to molecular systems appear in for diatomics and clusters, where degenerate orbitals are filled to maximize multiplicity, as seen in transition metal-doped superatoms like FeMg₈, yielding a magnetic moment of 4.0 μ_B and a HOMO-LUMO gap of 0.64 due to d-state hybridization. In complexes, such as those in organometallic , high-spin states guided by Hund's rules enhance ferromagnetic interactions, boosting reaction kinetics in processes like . Current research leverages these principles to predict magnetic properties, with enforced Hund's rules in DFT+U accurately forecasting easy-axis MA in rare-earth magnets (e.g., matching experimental directions for Dy in R₂Fe₁₄B), though amplitudes are overestimated for lighter elements like SmCo₅. In catalysis, Hund's-favored high-spin configurations in strongly correlated metals promote spin-polarized surfaces that lower activation barriers, as in fuel cells. For heavy elements, limitations arise from spin-orbit coupling, necessitating relativistic DFT to capture 4f localization; without it, standard DFT fails to reproduce Hund's ground states in actinides. The evolution from empirical heuristics to ab initio validation is exemplified by self-interaction-free relativistic local density approximations, which justify all three Hund's rules in solids like γ-Ce through orbital-dependent functionals, confirming high-spin, high-L, and positive J states without assumptions. This theoretical foundation underscores the rules' enduring utility, despite occasional exceptions in light atoms where correlation effects invert multiplicities.

References

  1. [1]
    Hund's rules - Oxford Reference
    These rules were put forward by the German physicist Friedrich Hund (1896–1993) in 1925. Hund's rules are explained by quantum theory involving the repulsion ...
  2. [2]
    Hund's Rules for Atomic Energy Levels - HyperPhysics Concepts
    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.
  3. [3]
    Hund's rules | Zeitschrift für Physik D Atoms,Molecules and Clusters
    We review the present state of our undertanding of Hund's first and second rules, their domains of validity, and of generalizations in cases where the.
  4. [4]
    Exploring Hund's rule with high-performance computing
    Nov 11, 2024 · For atoms like carbon, Hund's rule correctly predicts that the triplet state (with unpaired electrons) is more stable than the singlet state ( ...
  5. [5]
    Friedrich Hund: A Pioneer of Quantum Chemistry (1896–1997)
    Sep 24, 2022 · (i) 'Hund's rules' dealing with the electronic configuration of atoms and the atomic term symbols. These rules are now a part of high school ...
  6. [6]
    [PDF] Friedrich Hund: A Pioneer of Quantum Chemistry
    Sep 9, 2022 · Hund stated three rules for determining the ground state term symbol of an atom, in terms of the total electronic spin S, the total spatial ...
  7. [7]
    [PDF] Friedrich Hermann Hund - Resonance Sept 2022.cdr
    During 1925-1926, the fundamental laws of quantum mechanics were discovered, and it became established unusually rapidly as the fundamental branch of ...
  8. [8]
    Friedrich Hund and Chemistry - Kutzelnigg - 1996
    Apr 1, 1996 · Hund's first tow rules for atomic states have been found to be special cases of more general rules. Molecules can violate Hund's first rule ...Missing: original | Show results with:original
  9. [9]
    [PDF] Chemistry 21b – Spectroscopy Lecture # 14 – Electronic Structure ...
    Feb 5, 2018 · The physical basis of Hund's rule is that high spin states correlate with anti-symmetric spatial wavefunctions that, on average, have average ...Missing: mechanics | Show results with:mechanics
  10. [10]
    [PDF] SOLID STATE PHYSICS PART III Magnetic Properties of Solids - MIT
    The physical origin of the first Hund's rule is that to minimize the Coulomb repulsion between two electrons it is advantageous to keep them apart, and by ...
  11. [11]
    [PDF] Contents - Physics Courses
    ... explanation of Hund's first rule, which says to max- imize S. This raises an interesting point, because we know that the ground state spatial wavefunction.<|control11|><|separator|>
  12. [12]
    [PDF] Basics of the Theory of Atom
    Spin-orbit coupling and the Third Hund's Rule. The spin-orbit coupling is the relativistic effect giving rise to the so-called fine structure (small.
  13. [13]
    Zur Deutung der Molekelspektren. II | Zeitschrift für Physik A ...
    Hund, F. Zur Deutung der Molekelspektren. II. Z. Physik 42, 93–120 (1927). https://doi.org/10.1007/BF01397124. Download citation. Received: 07 February 1927.
  14. [14]
    Linienspektren und Periodisches System der Elemente - SpringerLink
    Dieser Buchtitel ist Teil des Digitalisierungsprojekts Springer Book Archives mit Publikationen, die seit den Anfängen des Verlags von 1842 erschienen sind.
  15. [15]
  16. [16]
    Atomic Data for Oxygen (O ) - Physical Measurement Laboratory
    Atomic Data for Oxygen (O) Atomic Number = 8 Atomic Weight = 15.9994 Reference E95 OI Ground State 1s 2 2s 2 2p 4 3 P 2 Ionization energy 109837.02 cm
  17. [17]
    Atomic Data for Carbon (C ) - Physical Measurement Laboratory
    C I Ground State 1s22s22p2 3P0. Ionization energy 90820.45 cm-1 (11.26030 eV) Ref. J66 C II Ground State 1s22s22p 2P°1/2. Ionization energy 196664.7 cm-1 ...
  18. [18]
    Atomic Data for Nitrogen (N ) - Physical Measurement Laboratory
    Nitrogen Neutral Atom Persistent Lines Nitrogen Neutral Atom Energy ... N II Ground State 1s22s22p2 3P0. Ionization energy 238750.3 cm-1 (29.6013 eV) ...Missing: spectra | Show results with:spectra
  19. [19]
    NIST Atomic Spectra Database Ionization Energies Data
    At. Num. El. name, Isoel. Seq. Ground Shells a, Ground Level, Ionization Energy (eV), Uncertainty (eV), References. 26, Iron, Fe, [Ar]3d64s2, 5D4 ...
  20. [20]
    [PDF] Spectroscopic Ground States - Dalal Institute
    Hence, all the 10 microstates for d1 and d9-configurations are distributed in 2D term symbols. 8. d2 and d8-configuration: For parallel arrangements, lz. −2. ↑.
  21. [21]
    A rigorous derivation of Hund's rules - Physics Stack Exchange
    Apr 11, 2023 · Hund's rule is the interplay between the Pauli exclusion principle and the Coulomb interaction, which was first pointed out by Heitler and ...Missing: old | Show results with:old
  22. [22]
    Energy Levels of Neutral Helium ( He I )
    Neutral Helium (He I) has energy levels such as 1s² at 0.000 cm⁻¹, 1s²s at 159855.9745 cm⁻¹, 1s²s at 166277.4403 cm⁻¹, and 1s²p at 169086.7666 cm⁻¹.
  23. [23]
    Persistent Lines of Neutral Sodium ( Na I )
    Sodium Neutral Atom Energy Levels Sodium Singly Ionized Persistent Lines ... 500 5895.924 0.614 0.000 3s 2S 1/2 R56 M03 16956.172 3p 2P° 1/2 60 8183.256 0.453 ...
  24. [24]
    [PDF] Atomic Terms,Hund's Rules, Atomic Spectroscopy
    May 6, 2008 · Atomic terms group similar energy quantum microstates. Hund's rules determine their energy. Atomic spectroscopy provides information on allowed ...
  25. [25]
  26. [26]
  27. [27]
    [PDF] The Spectrum and Energy Levels of the Neutral Atom of Boron (B1)
    The published data on the spectrum of the neutral atom of boron are compiled and presented. In one table. 164 lines in the range 36010-993 Å are listed with ...Missing: lower | Show results with:lower
  28. [28]
  29. [29]
    Importance of enforcing Hund's rules in density functional theory ...
    Jun 5, 2025 · Additionally, the Hund's rule states are enforced for all spin quantization directions in all rare-earth elements. Crystal structure.
  30. [30]
    [PDF] Importance of Enforcing Hunds Rules in Density Func - OSTI.GOV
    Overall, the 4f electron configurations in solids, especially those of heavy R elements, generally obey the same Hund's rules as in a free ion, according to the ...
  31. [31]
    A Density Functional Theory for the Average Electron Energy
    Jan 24, 2023 · (meta-GGAs) and hybrid functionals can be more realistic than those ... However, Hund's rule is never violated for single atoms in the considered ...
  32. [32]
    Hund's rule in superatoms with transition metal impurities - PNAS
    Jun 6, 2011 · Unlike atoms, the superatom orbitals span over multiple atoms and the filling of orbitals does not usually exhibit Hund's rule seen in atoms.
  33. [33]
    Strongly Correlated Electrons in Catalysis: Focus on Quantum ...
    Nov 10, 2021 · As a rule of thumb, reaction kinetics (thus catalytic activity) generally increase when interatomic ferromagnetic (FM) interactions are dominant ...
  34. [34]
    Review on Magnetism in Catalysis: From Theory to PEMFC ... - NIH
    In the present review, we introduce the relationship between magnetism and catalysis and outline its importance in the production of clean energy.<|separator|>
  35. [35]
  36. [36]
    Ab initio interpretation of Hund's rule for the methylene molecule ...
    Oct 10, 2007 · In the interpretation of Hund's rule there is an essential difference between atoms and molecules. For molecules one has to take account of ...