Fact-checked by Grok 2 weeks ago

Spdf

In and , spdf denotes the subshells of atomic orbitals distinguished by the ℓ, with s for ℓ=0 (spherical ), p for ℓ=1 (-shaped), d for ℓ=2 (cloverleaf or double ), and f for ℓ=3 (more complex geometries). These subshells organize electrons into principal levels (n), influencing atomic spectra, , and periodic properties through their and spatial distribution. The letters s, p, d, and f trace their origins to early 20th-century spectroscopic classifications of emission lines, labeled as (s), principal (p), diffuse (d), and fundamental (f) series based on their visual characteristics in spectra. This empirical naming, formalized by spectroscopists like Friedrich Paschen and later adopted in by , replaced numerical designations for secondary quantum numbers with these mnemonic labels. Each spdf subshell contains 2ℓ + 1 orbitals, yielding maximum electron capacities of 2 (s), 6 (p), 10 (d), and 14 (f), which dictate filling orders via the and explain valence shell behaviors across elements. Higher subshells (g, h, etc.) extend the scheme for ℓ > 3, though spdf dominate in naturally occurring elements up to . The notation underpins electron configuration predictions, such as [Ar] 4s² 3d¹⁰ 4p⁶ for , enabling causal insights into reactivity and magnetism without reliance on ad hoc adjustments.

Definition and Notation

Origins of the Letters

The letters s, p, d, and f designate atomic orbitals corresponding to azimuthal quantum numbers l = 0, 1, 2, and 3, respectively, and trace their origins to empirical classifications of series in emission spectra during the late 19th century. These terms—standing for sharp, principal, diffuse, and fundamental—arose from observations of line intensities, widths, and patterns, rather than any direct reference to orbital shapes or quantum properties. In 1872–1880, British spectroscopists George Liveing and analyzed spectra and categorized lines as "sharp" (narrow and well-defined), "principal" (prominent and intense, resembling 's main series), or "diffuse" (broader and less resolved), based on their visual appearance in spectrograms. By the 1880s, Balmer's for lines (1885) inspired extensions by Heinrich Kayser, , and , who formalized series as principal (p, strong absorptions akin to ), sharp (s, fine doublets or triplets), and diffuse (d, subordinate broader groups). Arno Bergmann identified the fourth "fundamental" (f) series in 1907 as the next subordinate progression, completing the quartet observed in heavier s. With the advent of , and others used these letters as shorthand for series constants (μ) in Rydberg's generalized equation for spectral frequencies. formalized their application to states in Max Born's 1925 Vorlesungen über Atommechanik, replacing numerical labels for the secondary with s, p, d, f to denote spectroscopic transitions. By the 1930s, this notation permeated quantum chemical descriptions of configurations, decoupling from spectral visuals while retaining the historical labels for subshells.

Usage in Electron Configurations

The spdf notation designates subshells in atomic electron configurations, where the principal quantum number n (1, 2, 3, ...) specifies the or , followed by a letter indicating the l (s for l=0, p for l=1, d for l=2, f for l=3), and a superscript denoting the number of electrons occupying that subshell. This format succinctly represents the distribution of s, as in the ground-state configuration of sodium ( 11): 1s² 2s² 2p⁶ 3s¹. Electrons occupy subshells according to the , filling from lowest to highest energy, with the order generally following 1s, 2s, 2p, 3s, 3p, 4s, , 4p, 5s, , 5p, , 4f, 5d, 6p, and so on, though exceptions occur due to electron-electron interactions, such as in (⁵ 4s¹ instead of ⁴ 4s²) to achieve a half-filled d subshell for greater stability. Each subshell has a maximum electron capacity of 2(2l + 1), yielding 2 electrons for s, 6 for p, 10 for d, and 14 for f, reflecting the number of available magnetic quantum numbers m_l (from -l to +l) times two spin states per orbital. In practice, configurations are derived by adding electrons sequentially to an atom's , adhering to the (no more than two electrons per orbital with opposite spins) and Hund's rule (maximizing unpaired electrons in degenerate orbitals for lowest energy). For ions, electrons are removed from the highest-energy subshell first, as in Fe²⁺ ([Ar] 3d⁶) from neutral iron ([Ar] 4s² 3d⁶). This notation enables prediction of chemical properties, such as valence electrons in outer subshells determining reactivity.
Subshelll ValueMax. ElectronsExample Elements
s02H (1s¹), He (1s²)
p16C (2p²), Ne (2p⁶)
d210Cr (3d⁵), Zn (3d¹⁰)
f314Ce (4f¹ in [Xe] 6s² 5d¹ 4f¹), Lu (4f¹⁴)

Quantum Mechanical Basis

Azimuthal Quantum Number

The , denoted \ell, specifies the subshell within a given principal energy level n and characterizes the orbital of an in . It takes nonnegative integer values ranging from to n-1, thereby determining the number of subshells available for a principal quantum number n; for instance, n=1 allows only \ell=0, while n=3 permits \ell=0,1,2. This quantum number arises naturally from the time-independent for the , solved in spherical coordinates, where the angular portion of the wave function separates into \Theta(\theta) and \Phi(\phi) functions. The eigenvalue \ell(\ell+1)\hbar^2 governs the centrifugal term in the of the radial , influencing the orbital's radial and . Specific values of \ell define distinct orbital types: \ell=0 yields spherical s-orbitals with no angular nodes; \ell=1 produces dumbbell-shaped p-orbitals with one angular nodal ; \ell=2 corresponds to more d-orbitals with two nodal planes; and \ell=3 to f-orbitals with three. The P_\ell^{|m|}(\cos\theta) in the \Theta impose conditions that quantize \ell, ensuring the wave remains finite and single-valued over . In multi-electron atoms, \ell retains its role in labeling subshells, though electron-electron interactions modify energies via screening and effects, with higher \ell subshells generally higher in energy within the same n due to poorer to the . The orbital vector has magnitude \sqrt{\ell(\ell+1)}\hbar and z-component m_\ell \hbar (where m_\ell = -\ell, \dots, +\ell), linking \ell to the degeneracy of subshells: $2\ell + 1 orbitals per subshell. This framework, derived from first solving the radial equation independently of \ell for (where energy depends only on n), extends to approximate solutions for heavier atoms using Hartree-Fock methods or .

Relationship to Angular Momentum

The azimuthal quantum number l, which labels the spdf subshells as l = 0 (s), l = 1 (p), l = 2 (d), and l = 3 (f), determines the quantized orbital of an in an . This ranges from 0 to n-1, where n is the , and its value for a given subshell fixes the magnitude and possible projections of the vector \mathbf{L}. The magnitude of \mathbf{[L](/page/L')} is |\mathbf{L}| = \sqrt{l(l+1)} \hbar, where \hbar = h / 2\pi and h is Planck's , rather than the naive classical l \hbar. This form emerges from the eigenvalues of the squared angular momentum operator \hat{L}^2, which satisfy \hat{L}^2 \psi = l(l+1) \hbar^2 \psi for eigenfunctions \psi described by Y_l^{m_l}(\theta, \phi), reflecting the inherent quantum uncertainty in the direction of \mathbf{L} perpendicular to any chosen axis. For s subshells (l = 0), |\mathbf{L}| = 0, corresponding to no orbital and spherically symmetric probability . In p subshells (l = 1), |\mathbf{L}| = \sqrt{2} \hbar \approx 1.414 \hbar, with three orbitals arising from magnetic quantum numbers m_l = -1, 0, +1, yielding z-component projections m_l \hbar. For d subshells (l = 2), |\mathbf{L}| = \sqrt{6} \hbar \approx 2.449 \hbar and five orbitals ($2l + 1 = 5); f subshells (l = 3) yield |\mathbf{L}| = \sqrt{12} \hbar \approx 3.464 \hbar and seven orbitals. The number of orbitals in a subshell, given by $2l + 1, equals the number of possible m_l values, each specifying a distinct of the angular momentum vector relative to an external , as observed in Zeeman splitting of lines. Thus, the spdf designation encodes the electron's potential for orbital , which affects transition probabilities in atomic spectra and in multi-electron atoms.

Orbital Characteristics

Shapes and Orientations

The shapes of atomic orbitals are governed by the l, which dictates the angular part of the wave function and thus the orbital's . For l = 0 (s orbitals), the shape is spherically symmetric, with electron probability density highest near the and decreasing radially without nodes. This symmetry arises from the absence of , resulting in a single orbital per subshell that is independent of direction. For l = 1 (p orbitals), the shape features two lobes of opposite phase separated by a nodal plane through the , resembling a aligned along a principal axis. These orbitals possess one angular nodal plane, concentrating probability density along the axis perpendicular to the . In l = 2 (d orbitals), shapes become more intricate, typically with two angular nodal planes: common forms include four-lobed cloverleaf patterns in the xy-plane or double- configurations with lobes along z and perpendicular rings. These exhibit higher , leading to two nodal surfaces that divide space into distinct regions of probability. For l = 3 (f orbitals), shapes are highly complex, featuring three angular nodal planes and often visualized as combinations of lobes and tori, with probability densities separated by multiple s that reflect the increased . The number of angular nodes equals l, increasing structural complexity and directional preferences as l rises. Orbital orientations within a subshell are specified by the m_l, which ranges from -[l](/page/L') to +[l](/page/L') in integer steps, yielding $2l + 1 possible values and thus degenerate orbitals differing only in spatial alignment. For s orbitals (l = 0), m_l = 0 only, conferring no distinct orientation due to . p orbitals (l = 1) have m_l = -1, 0, +1, corresponding to orientations along the y, z, and x axes, respectively, in a Cartesian , with each to the others for maximal separation. d orbitals (l = 2) offer five orientations (m_l = -2 to +2), including planar quadrupolar (m_l = \pm 2) and axial (m_l = 0) forms, while f orbitals (l = 3) provide seven, with alignments involving higher-order that project lobes in varied azimuthal and polar directions. In the absence of an external , these orientations are energetically equivalent, but m_l quantizes the z-component of as m_l \hbar.

Energy Levels and Filling Order

In multi-electron atoms, the energy of a subshell is determined by both the principal quantum number n and the azimuthal quantum number l, with energies increasing as n rises and, for fixed n, increasing from s (l=0) to p (l=1) to d (l=2) to f (l=3). This l-dependence stems from differences in radial electron probability distributions: s subshells penetrate closer to the nucleus on average, resulting in less shielding by inner electrons and a higher effective nuclear charge, whereas p, d, and f subshells have nodal structures that keep electron density farther out, increasing their energy relative to s for the same n. The order in which subshells are filled with electrons follows the , prioritizing orbitals of lowest energy to minimize total energy in the . This filling sequence is predicted by the Madelung energy ordering rule, which arranges subshells by ascending n + l; for subshells with equal n + l, the one with smaller n fills first. The resulting order is 1s, 2s, 2p, 3s, 3p, 4s, 3d, 4p, 5s, 4d, 5p, 6s, 4f, 5d, 6p, 7s, 5f, 6d, 7p. Exceptions occur when alternative configurations lower energy through exchange stabilization or half/full subshell preferences, as in (3d⁵4s¹ instead of 3d⁴4s²) and (3d¹⁰4s¹ instead of 3d⁹4s²), where d subshell completion outweighs strict n + l ordering. Similar irregularities appear for and 5d in lanthanides and actinides, though the general Aufbau pattern holds for predicting configurations up to 118.

Historical Development

Spectroscopic Line Classification

The spectroscopic notation s, p, d, and f originated in the late 19th and early 20th centuries from the empirical classification of emission line series in the spectra of alkali metals, such as sodium and potassium, observed under high-resolution spectroscopy. Spectroscopists, including Johannes Rydberg and others, identified recurring patterns of lines grouped into four main series—principal, sharp, diffuse, and fundamental—based on their wavelengths, intensities, and visual characteristics like sharpness or diffuseness, which arose from factors such as fine structure splitting and resolution limits of early instruments. The principal series (p) comprised the most intense and prominent lines, typically resulting from transitions between states with principal quantum number differences, often from higher p-designated terms to the ground s state in atoms. The sharp series (s) featured narrow, precisely resolved lines due to minimal splitting, corresponding to transitions from excited s terms to a fixed p term. In contrast, the diffuse series (d) exhibited broader, less resolved lines attributed to greater effects causing increased splitting, involving d terms transitioning to p states. The fundamental series (f), weaker and less prominent, involved f terms and completed the classification for higher states. These labels were initially applied to spectroscopic terms (e.g., ^2S, ^2P) denoting levels rather than orbitals, with the series names reflecting observational priorities: "principal" for dominance in spectra akin to hydrogen's main lines, "sharp" for clarity, "diffuse" for haziness, and "fundamental" as a category. Rydberg's work in the formalized the series using empirical formulas for wavelengths, such as extensions of the Balmer-Rydberg , enabling prediction of unobserved lines and highlighting regularities across . By 1914, Alfred Fowler and others had mapped these series to multi-electron transitions, laying groundwork for notation without invoking atomic structure theories. This classification system proved robust, accommodating data from improved grating spectrometers that revealed thousands of lines, but relied on phenomenological descriptions rather than causal mechanisms, as was absent until the . Empirical verification came from vapor discharge tubes, where line positions matched Rydberg constants adjusted for , with principal series lines dominating visible spectra (e.g., sodium D-lines at 589.0 nm and 589.6 nm from 3p to 3s). Subsequent extensions to g, h, etc., followed analogous patterns for rarer spectra, though s–f remained standard for common elements.

Transition to Quantum Theory

In the old quantum theory, the empirical spectroscopic series classifications were theoretically rationalized through extensions of Bohr's 1913 atomic model. , in , refined the model by permitting elliptical orbits to account for relativistic and spectral anomalies in alkali metals, introducing a subsidiary quantum number k (ranging from 1 to the principal quantum number n) alongside Bohr's n. This k quantized the component of , with the quantization condition J_k = k h / 2\pi, and directly corresponded to the observed spectral series: k=1 for sharp (s) lines, characterized by minimal splitting; k=2 for principal (p) series; k=3 for diffuse (d); and k=4 for fundamental (f). Sommerfeld's framework reproduced the Rydberg formulas for multiple series and explained selection rules like \Delta k = \pm 1, bridging empirical observations to quantized dynamics without full wave mechanics. Sommerfeld's 1919 Atombau und Spektrallinien formalized these associations, demonstrating how intra-atomic electric fields and quantized orbits generated the hierarchical series structure seen in spectra, such as the sodium D-line . This semi-classical approach marked the initial theoretical adoption of spdf designations, shifting from empirical labeling to a quantized orbital framework, though it still relied on classical trajectories. The model's success in predicting spectra energies, within experimental precision of about 0.1%, validated the linkage, but limitations emerged in multi-electron systems and anomalous Zeeman effects. The full transition to modern quantum mechanics occurred in the mid-1920s with matrix mechanics (Heisenberg, Born, Jordan, 1925) and wave mechanics (Schrödinger, 1926), where the subsidiary number evolved into the l = k - 1. Solutions to the separated into radial and angular parts, with l governing Y_l^m(\theta, \phi), determining orbital shapes and \sqrt{l(l+1)} \hbar. The spdf labels persisted for l=0 (s, spherical, no angular nodes), l=1 (p, dumbbell), l=2 (d, cloverleaf), and l=3 (f, complex), as they aligned with spectroscopic intensities and dipole selection rules \Delta l = \pm 1. This quantum mechanical interpretation confirmed the old theory's mappings while resolving inconsistencies, such as precise multi-electron interactions via Pauli exclusion (1925). By 1927, integrated spdf notation into atomic and molecular electron configurations, replacing numerical k or l values to denote subshells in the periodic table, as detailed in his Linienspektren und Periodensystem. This adoption facilitated building configurations like 1s² 2s² 2p⁶ for , matching empirical ionization potentials and spectral data. Hund's work, building on Born's refinements to Bohr's erroneous 1922 subshell assignments, entrenched the notation in , with widespread use in literature by the 1930s, such as Rabinowitsch and Thilo's 1930 . The persistence of spdf reflected its empirical fidelity now underpinned by causal quantum principles, rather than mere convention.

Applications and Extensions

Beyond f-Orbitals

Higher angular momentum orbitals, corresponding to azimuthal quantum numbers ℓ ≥ 4, extend the quantum mechanical description of atomic electron states beyond the f subshell (ℓ = 3). These include g orbitals (ℓ = 4), h orbitals (ℓ = 5), i orbitals (ℓ = 6), and subsequent letters, with the sequence skipping 'j' to avoid confusion with the . For a given n, ℓ ranges from 0 to n-1, permitting g orbitals starting at n=5 and higher subshells at larger n; the wavefunctions derive from Y_ℓ^m(θ, φ), yielding increasingly intricate angular distributions with ℓ nodal surfaces. In multi-electron atoms, these orbitals remain unoccupied in ground-state configurations through element 118 (), as electron filling adheres to the Madelung rule, prioritizing lower n and ℓ up to 7p. Theoretical models predict g subshell involvement beyond Z=120, potentially stabilizing new blocks in an , though relativistic effects—such as orbital contraction and spin-orbit splitting—destabilize such configurations in superheavy nuclei with lifetimes under microseconds. computations, using methods like Dirac-Fock, incorporate g and h contributions for accurate energies in heavy elements like (Z=103), where admixtures affect ionization potentials by up to 10%. Observationally, higher ℓ orbitals manifest in Rydberg states of atoms, where electrons occupy high-n levels (n > 20) with ℓ ≈ n-1, forming near-circular orbits minimally penetrating the core; spectroscopy has resolved 5g states in (n=19–30) with binding energies matching hydrogenic predictions adjusted for quantum defects of order 0.01 eV. These states enable applications in Rydberg blockade for quantum , achieving fidelities exceeding 99% in trapped experiments, and in precision measurements of atomic polarizabilities. In , g-like orbitals influence beta decay rates in transactinides, as modeled in shell corrections to barriers.
Subshellℓ ValueNumber of Orbitals (2ℓ+1)Minimum n for Occupancy
g495
h5116
i6137
Such tabulations highlight the degeneracy and radial extent, with higher ℓ orbitals exhibiting reduced and thus weaker interactions, quantified by values ⟨r⟩ scaling as n² / Z_eff. While empirically unpopulated in , their inclusion refines predictions for molecular properties in exotic systems, like clusters.

Role in Periodic Table and Chemistry

The periodic table's block structure reflects the sequential filling of atomic subshells defined by the l: the s-block (l = 0) encompasses groups 1 and 2, where valence electrons occupy ns orbitals; the p-block (l = 1) spans groups 13–18 with np valence electrons; the d-block (l = 2) covers groups 3–12, involving (n-1)d subshells; and the f-block (l = 3) includes the lanthanides and actinides with (n-2)f electrons. This division stems from the , which prioritizes lower-energy orbitals based on increasing n and l values, resulting in 2, 6, 10, and 14 electrons per respective block (excluding s pairing). Subshell-specific electron configurations dictate periodic trends in atomic properties. For instance, s- and p-block elements exhibit valence electrons in more penetrating orbitals, leading to smoother decreases in atomic radius across periods and predictable ionization energies, whereas d- and f-block electrons provide poorer shielding due to their nodal structures and radial distribution, causing contractions like the lanthanide contraction (atomic radii decrease by ~10–15 pm from La to Lu despite increasing n). This shielding inefficiency elevates effective nuclear charge for subsequent elements, enhancing electronegativities and influencing trends such as higher ionization energies in post-transition metals compared to expectations from n alone. In chemical reactivity, subshell occupancy governs bonding and electronic transitions. s-Block metals form highly electropositive ions (e.g., Na⁺ from 3s¹) due to low ionization energies (~5.1 eV for Na), driving ionic compounds and reducing properties. p-Block elements leverage directional p orbitals for covalent bonding, yielding diverse oxidation states (e.g., C from +4 to -4). d-Block transition metals display variable valences (up to 10 electrons available) and paramagnetism from unpaired d electrons, enabling coordination complexes and d-d absorption spectra in the visible range (e.g., [Ti(H₂O)₆]³⁺ at 500 nm). f-Block elements feature localized 4f/5f electrons with minimal hybridization, resulting in predominantly +3 oxidation states, high coordination numbers (>6), and strong magnetic moments (e.g., Gd³⁺ with 7 unpaired electrons, μ ≈ 7.9 μ_B). These traits underpin applications like catalysis in d-block alloys and fission in actinides.

References

  1. [1]
    The periodic table, electron shells, and orbitals - Khan Academy
    We can break each electron shell down into one or more subshells, which are simply sets of one or more orbitals. Subshells are designated by the letters ‍ , ‍ ...
  2. [2]
    S P D F Orbitals and Angular Momentum Quantum Numbers
    Jun 9, 2025 · The orbital names s, p, d, and f stand for names given to groups of lines originally noted in the spectra of the alkali metals. These line ...<|separator|>
  3. [3]
    What is the historical reasoning for electron orbital names?
    Dec 12, 2014 · The basis is actually German, not Latinate. They stand for: scharf (sharp), prinzipal (principal), diffus (diffuse), and fundamental (fundamental).
  4. [4]
    Shells, subshells, and orbitals (video) - Khan Academy
    Sep 6, 2019 · What it seems like is shells are the energy levels, subshells are s, p, d, and f, and orbitals seem to be each individual form of each of the s p d f subshells.
  5. [5]
    Electron Configurations - FSU Chemistry & Biochemistry
    The symbols used for writing the electron configuration start with the shell number (n) followed by the type of orbital and finally the superscript indicates ...Missing: usage | Show results with:usage
  6. [6]
    [PDF] Chapter 6 Electronic Structure of Atoms - MSU chemistry
    Azimuthal Quantum Number, l l = 0, 1...,n-1. Value of l. 0 1 2 3. Type of orbital. s p d f ... ➢ Letter denoting the type of orbital. ➢ Superscript denoting ...
  7. [7]
    8.3: Electron Configurations- How Electrons Occupy Orbitals
    Aug 14, 2020 · The relative energy of the subshells determine the order in which atomic orbitals are filled (1s, 2s, 2p, 3s, 3p, 4s, 3d, 4p, and so on).
  8. [8]
    Quantum Numbers and Electron Configurations
    These quantum numbers describe the size, shape, and orientation in space of the orbitals on an atom.
  9. [9]
    Electron configurations article - Khan Academy
    Electron configurations are a simple way of writing down the locations of all of the electrons in an atom. As we know, the positively-charged protons in the ...Missing: usage | Show results with:usage
  10. [10]
    Quantum Numbers - Specifying the electron state - Grandinetti Group
    ℓ : azimuthal quantum number. The azimuthal quantum number has integral values of ℓ = 0 to ℓ = n - 1 for each value of n. This quantum number defines the ...
  11. [11]
    Hydrogen Schrodinger Equation - HyperPhysics
    The Azimuthal Equation​​ This is the origin of the magnetic quantum number. The expression of the separation constant in terms of this quantum number affects the ...
  12. [12]
    3.2: Quantum Numbers for Atomic Orbitals - Chemistry LibreTexts
    Aug 13, 2021 · Schrödinger's approach uses three quantum numbers (n, l, and ml) to specify any wavefunction. The quantum numbers provide information about ...Quantum Numbers · The Magnetic Quantum Number
  13. [13]
    1.2: Quantum Numbers - Chemistry LibreTexts
    Aug 27, 2023 · Quantum mechanics uses four quantum numbers (n, l, ml, and ms) to define wavefunction ... The Angular Momentum (or Azimuthal) Quantum Number.
  14. [14]
    6.6: Orbital Angular Momentum and the p-Orbitals
    Mar 8, 2025 · This page discusses the relationship between classical and quantum angular momentum for electrons in atoms, emphasizing quantum numbers ...
  15. [15]
    2.2: Atomic Orbitals and Quantum Numbers - Chemistry LibreTexts
    Jun 5, 2019 · The letter in the orbital name defines the subshell with a specific ... angular momentum or azimuthal quantum number, l, 0 ≤ l ≤ n – 1 ...Missing: correspondence | Show results with:correspondence
  16. [16]
    Quantum Numbers for Atoms - Chemistry LibreTexts
    Aug 14, 2024 · The Orbital Angular Momentum Quantum Number ( l ) · determines the shape of an orbital, and therefore the angular distribution. ·. · indicates a ...Missing: spdf | Show results with:spdf
  17. [17]
    [PDF] Chapter 17 Many-Electron Atoms and Chemical Bonding
    The energy of an orbital of a multielectron atom depends on n and l (but not m l. ) ... The 4s and 3d orbitals are close in energy in the one electron atom.
  18. [18]
    Energy of orbitals - BYJU'S
    Nov 1, 2020 · Thus, the energy of orbitals depends upon the values of both the principal quantum number (n) and the azimuthal quantum number (l). Hence, the ...<|control11|><|separator|>
  19. [19]
    Multi-electron Atoms - Chemistry LibreTexts
    Jan 29, 2023 · Energy levels consist of orbitals and sub-orbitals. The lower the energy level the electron is located at, the closer it is to nucleus. As we go ...
  20. [20]
    Electron Configuration - Chemistry LibreTexts
    Jan 29, 2023 · The electron configuration is the standard notation used to describe the electronic structure of an atom.
  21. [21]
    The Order of Filling 3d and 4s Orbitals - Chemistry LibreTexts
    Jan 29, 2023 · The aufbau principle explains how electrons fill low energy orbitals (closer to the nucleus) before they fill higher energy ones.
  22. [22]
    The Origin of the s, p, d, f Orbital Labels - ResearchGate
    Aug 6, 2025 · This is the origin of the familiar modern description of s, p, d, and f (etc.) applied to atomic orbitals.
  23. [23]
    [PDF] Spectroscopic Notation - Macmillan Learning
    The sharp series corresponded to transitions from the higher sharp (S terms) energy states—the running terms—to the lowest principal (P term) state, the ...<|separator|>
  24. [24]
    The Spectral Lines of Hydrogen - Spectroscopy Online
    Nov 1, 2008 · Bohr's derivation of Rydberg's formula is presented in Appendix 2. For his groundbreaking work, which led to the development of modern ...
  25. [25]
    26. The Sommerfield Model, Introduction of the Quantum Numbers.
    Sep 28, 2016 · Sommerfeld replaced Bohr's circular orbits with elliptical orbitals or shells and introduced azimuthal and magnetic quantum numbers. He stated ...
  26. [26]
    Atomic structure and spectral lines : Sommerfeld, Arnold, 1868-1951
    Jul 27, 2019 · Atomic structure and spectral lines ; Publication date: 1923 ; Topics: Atoms, Matter -- Constitution, Spectrum analysis ; Publisher: New York, ...
  27. [27]
    Old Quantum Mechanics by Bohr and Sommerfeld from a Modern ...
    Jun 30, 2025 · We review Bohr's atomic model and its extension by Sommerfeld from a mathematical perspective of wave mechanics. The derivation of quantization ...
  28. [28]
    [PDF] The Origin of the s, p, d, f Orbital Labels - UC Homepages
    Beginning in the 1930s both. Hund's corrected configurations and his s, p, d, f nota- tion began to slowly leak into the chemical literature, where they have ...
  29. [29]
    Q: Is it possible for an atomic orbital to exist beyond the s, p, f and d ...
    Jan 4, 2012 · There's no reason for electrons not to fill sub-shells past 'f', but it's not needed until element 121, which would need a 'g' orbital.Missing: ghi | Show results with:ghi
  30. [30]
    The Orbitron: 6g atomic orbitals - Mark Winter
    There are nine 6g orbitals. The higher g-orbitals (7g) are more complex since they have more spherical nodes while lower g-orbitals (6g) have none.
  31. [31]
    2.4 Electron Configurations - Chemistry LibreTexts
    Jun 5, 2019 · spdf Notation. The most common way to describe electron configurations is to write distributions in the spdf notation. Although the ...Exceptions · Writing Electron Configurations · Orbital Diagrams · spdf Notation
  32. [32]
    6.6: Electron Configurations and the Periodic Table
    After the 4s subshell is filled, the 3d subshell is filled with up to 10 electrons. This explains the section of 10 elements in the middle of the periodic table ...
  33. [33]
    21.1 Overview of Periodic Trends
    The ionization energies increase because filled (n − 1)d or (n − 2)f subshells are relatively poor at shielding electrons in ns orbitals. Thus the two electrons ...
  34. [34]
    [PDF] Electronic Structure and Periodic Properties of Elements
    Figure 6.28 This periodic table shows the electron configuration for each subshell. By “building up” from hydrogen, this table can be used to determine the ...
  35. [35]
    Property Trends in the Periodic Table - Chemistry 301
    Cu, Ag, and Au are all fairly large as adding one electron will fill the d-subshell. This makes the trend left to right on the periodic table spotty at best.