Fact-checked by Grok 2 weeks ago

Aufbau principle

The Aufbau principle, from the word Aufbau meaning "building up," is a fundamental concept in that dictates the order in which s fill atomic orbitals in the of multi- atoms, starting from the lowest level and proceeding to higher ones. This principle enables the prediction of electron configurations by assuming s are added one at a time to orbitals of increasing , influenced by both the principal (n) and the (l). In practice, the Aufbau principle follows a specific sequence for orbital filling: 1s, 2s, 2p, 3s, 3p, 4s, 3d, 4p, 5s, 4d, 5p, 6s, 4f, 5d, 6p, 7s, 5f, 6d, and 7p, reflecting the relative energies of subshells in multi-electron atoms where penetration effects and electron shielding alter the simple hydrogen-like order. Each orbital can hold a maximum of two electrons with opposite spins, as dictated by the , ensuring no two electrons in an atom share identical quantum numbers. Complementing this, Hund's rule specifies that for degenerate orbitals (those of equal energy within a subshell), electrons occupy each orbital singly with parallel spins before pairing up, maximizing the atom's total spin multiplicity. While the Aufbau principle provides a reliable framework for most elements, exceptions occur in transition metals and heavier elements due to subtle differences, such as the of half-filled or fully filled subshells (e.g., chromium's 3d⁵ 4s¹ instead of 3d⁴ 4s²). These rules collectively underpin the of the periodic table, correlating electron configurations with elemental properties like , , and chemical reactivity.

Fundamentals of Electron Configuration

Definition and Core Principles

The Aufbau principle, derived from the German word for "building up," states that electrons in the of a multi-electron occupy orbitals in the order of increasing , starting with the lowest-energy orbital available and proceeding until all electrons are placed. This guideline provides a systematic approach to determining the , which describes how electrons are distributed among the orbitals to achieve the lowest possible total for the . The principle assumes that electrons are added sequentially to the , mimicking the construction of elements in the periodic table from onward. Central to the Aufbau principle are its integration with two fundamental quantum mechanical rules: the and Hund's rule. The stipulates that no two electrons within an atom can occupy the same simultaneously, limiting each orbital to a maximum of two electrons that must have opposite spins. Hund's rule complements this by directing that, for orbitals of equal energy (degenerate orbitals), electrons will first occupy separate orbitals with parallel spins before pairing up, thereby minimizing electron-electron repulsion and maximizing total spin for stability. Together, these principles ensure that the resulting reflects the atom's stable , influencing its chemical behavior and periodic properties. Atomic orbitals, the building blocks in this process, are probability distributions describing electron positions and are categorized by their subshell types—s, p, d, and f—each corresponding to different shapes and angular momenta. The s orbital is spherically symmetric around the , accommodating up to two electrons; p orbitals have a dumbbell shape with three orientations along the axes, holding up to six electrons; d orbitals feature more cloverleaf-like patterns with five orientations, up to ten electrons; and f orbitals are even more complex with seven orientations, up to fourteen electrons. These subshells are organized by n (indicating energy level) and l (defining subshell type, where l = 0 for s, 1 for p, 2 for d, and 3 for f). A straightforward illustration of the Aufbau principle is the of carbon (Z = 6), which has six electrons filling as 1s² 2s² 2p²: the first two electrons pair in the 1s orbital, the next two fill the 2s orbital, and the final two occupy separate 2p orbitals singly per Hund's rule, without delving into higher subshells. This configuration exemplifies how the principle prioritizes lower-energy orbitals to form the atom's overall structure, serving as the foundation for predicting properties across the periodic table.

Orbital Energies and Quantum Numbers

The atomic orbitals in quantum mechanics are defined by a set of four quantum numbers that uniquely specify the state of an electron in an atom. The principal quantum number n is a positive integer (n = 1, 2, 3, \dots) that primarily determines the size and energy of the orbital, with higher values corresponding to larger orbitals and higher energy levels. The azimuthal quantum number l, also known as the angular momentum quantum number, takes integer values from 0 to n-1 and defines the shape and subshell of the orbital, ranging from s (l=0) to higher subshells as l increases. The magnetic quantum number m_l specifies the orientation of the orbital in space and ranges from -l to +l in integer steps, allowing for $2l+1 possible orientations per subshell. Finally, the spin quantum number m_s describes the intrinsic spin of the electron, with possible values of +\frac{1}{2} or -\frac{1}{2}, ensuring that no two electrons in the same atom can have identical sets of all four quantum numbers, as per the Pauli exclusion principle. Together, these quantum numbers limit each orbital to a maximum capacity of two electrons with opposite spins. In hydrogen-like atoms (one-electron systems such as H or He^+), the energy of an electron depends solely on the principal quantum number n, with the energy levels given approximately by E_n = -\frac{13.6 \, \text{eV} \cdot Z^2}{n^2}, where Z is the atomic number; this degeneracy means orbitals with the same n but different l have identical energies. However, in multi-electron atoms, electron-electron repulsions introduce complexity, causing the energy to depend on both n and l: higher n generally increases energy due to larger orbital size, while for a fixed n, energy rises with l because of differences in electron-nucleus interactions. This dependence arises from the effective nuclear charge Z_{\eff}, which an electron experiences after accounting for shielding by inner electrons; Z_{\eff} is less than the full nuclear charge Z and varies with orbital type, leading to the ordering where, for example, 3s orbitals are lower in energy than 3p orbitals./Quantum_Mechanics/10:_Multi-electron_Atoms/Multi-Electron_Atoms/Penetration_and_Shielding) The l also classifies orbitals into types based on their and approximate shapes: s orbitals (l=0) are spherical with zero , p orbitals (l=1) are dumbbell-shaped along principal axes with \hbar \sqrt{l(l+1)} = \hbar \sqrt{2}, d orbitals (l=2) exhibit cloverleaf patterns in multiple planes with higher \hbar \sqrt{6}, and f orbitals (l=3) have more complex, multi-lobed shapes with \hbar \sqrt{12}./07:_Atomic_Structure_and_Periodicity/12.09:_Orbital_Shapes_and_Energies) These shapes reflect the probability distribution of the electron, derived from solutions to the , without implying rigid paths but rather regions of high . A key factor in the energy ordering within the same principal shell is the penetration effect, where s electrons (l=0) have a higher probability density near the nucleus compared to p (l=1), d (l=2), or f (l=3) electrons, allowing them to experience less shielding from inner electrons and thus a higher Z_{\eff}./Quantum_Mechanics/10:_Multi-electron_Atoms/Multi-Electron_Atoms/Penetration_and_Shielding) Shielding occurs as inner electrons reduce the net attraction felt by outer electrons, but penetration varies by subshell: s orbitals penetrate most effectively, followed by p, d, and f, resulting in progressively lower Z_{\eff} and higher energies for higher l. Qualitatively, the orbital energy can be approximated as E \propto -\frac{Z_{\eff}^2}{n^2}, where Z_{\eff} is estimated using methods like Slater's rules, which account for shielding constants that differ by subshell (e.g., s and p electrons shield less effectively than d or f). This variation in Z_{\eff} with l underlies the energy hierarchy essential for understanding electron arrangements in atoms.

The Madelung Ordering Rule

Derivation of the n + l Rule

The Madelung rule, also known as the n + l rule, states that atomic orbitals in multi-electron atoms are filled in the order of increasing , where is the principal and is the ; for orbitals with equal n + l, the one with smaller n is filled first. This rule arises from approximations in solving the Hartree-Fock equations for multi-electron atoms, where the single-particle orbital energies depend on both the principal and azimuthal quantum numbers due to the shaped by electron-electron repulsion. In these approximations, such as those using Slater-type orbitals or , the energy scales primarily with n + l because the principal quantum number n governs the average radial extent of the orbital, while the azimuthal quantum number l influences the centrifugal barrier and penetration toward the , balancing the kinetic and contributions. Specifically, higher l values lead to poorer shielding by inner electrons for a given n, raising the , but the combined n + l provides a simple index that captures the dominant ordering trend in neutral atoms. The resulting orbital energy ordering follows increasing n + l values, as illustrated in the table below for the initial subshells:
Orbitalnln + l
1s101
2s202
2p213
303
3p314
4s404
325
For cases of equal n + l, such as 2p and (both 3) or 3p and 4s (both 4), the lower n orbital has the lower energy. The n + l ordering is corroborated by atomic spectroscopy, where term symbols and ionization energies from spectral lines indicate that the 4s orbital lies below 3d for elements like potassium through zinc, reflecting the actual energy hierarchy in ground-state configurations. Despite its utility, the rule remains empirical rather than rigorously derived from first principles, as precise orbital energies are modulated by complex electron-electron interactions that introduce deviations, particularly in heavier atoms.

Orbital Filling Sequence

The orbital filling sequence in the Aufbau principle arranges subshells in order of increasing energy, primarily following the Madelung rule where subshells are ordered by the sum of the principal n and the (n + ℓ), with ties broken by increasing n. This results in the standard sequence: 1s → 2s → 2p → 3s → 3p → 4s → 3d → 4p → 5s → 4d → 5p → 6s → 4f → 5d → 6p → 7s → 5f → 6d → 7p. A common mnemonic for remembering this order is the "diagonal rule," visualized as a diagram with subshells arranged in rows by n (1s at the top, increasing downward) and columns by subshell type (s, p, d, f), where filling follows diagonal arrows from upper right to lower left, starting at 1s and progressing through the sequence. Each subshell has a fixed maximum electron capacity determined by the number of orbitals it contains, with each orbital holding up to two electrons of opposite : s subshells hold 2 electrons, p subshells hold 6, d subshells hold 10, and f subshells hold 14. This filling sequence corresponds to the blocks of the periodic table, where elements are grouped by the subshell being filled in their valence electrons: the s-block includes groups 1 and 2 ( and ), the p-block includes groups 13 through 18 (main group ), the d-block includes groups 3 through 12 (transition metals), and the f-block encompasses the lanthanides (4f) and actinides (5f). To illustrate the application, the electron configurations for the first 20 follow this strictly, building cumulatively until the 4s subshell begins after the 3p. Representative examples include:
Atomic NumberElementElectron Configuration
1H1s¹
2He1s²
3Li[He] 2s¹
4Be[He] 2s²
5B[He] 2s² 2p¹
6C[He] 2s² 2p²
7N[He] 2s² 2p³
8O[He] 2s² 2p⁴
9F[He] 2s² 2p⁵
10Ne[He] 2s² 2p⁶
11Na[Ne] 3s¹
12Mg[Ne] 3s²
13Al[Ne] 3s² 3p¹
14Si[Ne] 3s² 3p²
15P[Ne] 3s² 3p³
16S[Ne] 3s² 3p⁴
17Cl[Ne] 3s² 3p⁵
18Ar[Ne] 3s² 3p⁶
19K[Ar] 4s¹
20Ca[Ar] 4s²

Exceptions to the Rule

Anomalies in Transition Metals (d-block)

The Aufbau principle generally predicts that the 4s orbital fills before the 3d orbitals in the first transition series, leading to configurations like [Ar] 4s² 3dⁿ for scandium to manganese and [Ar] 4s² 3d^{n+1} thereafter. However, several d-block elements deviate from this order, adopting configurations that prioritize the stability of half-filled (d⁵) or fully filled (d¹⁰) subshells over the expected filling sequence. These anomalies arise because the energy difference between the ns and (n-1)d orbitals is small, allowing the exchange energy gained from symmetric electron distributions to outweigh the slight promotion cost from s to d orbitals. Common exceptions include (Cr, Z=24), which has the configuration [Ar] 4s¹ 3d⁵ instead of the predicted [Ar] 4s² 3d⁴; (Cu, Z=29), [Ar] 4s¹ 3d¹⁰ rather than [Ar] 4s² 3d⁹; (Nb, Z=41), [Kr] 5s¹ 4d⁴ instead of [Kr] 5s² 4d³; (Mo, Z=42), [Kr] 5s¹ 4d⁵ over [Kr] 5s² 4d⁴; (Ru, Z=44), [Kr] 5s¹ 4d⁷; (Rh, Z=45), [Kr] 5s¹ 4d⁸; (Pd, Z=46), [Kr] 4d¹⁰ (with empty 5s); silver (Ag, Z=47), [Kr] 5s¹ 4d¹⁰ instead of [Kr] 5s² 4d⁹; and (Au, Z=79), [Xe] 4f¹⁴ 5d¹⁰ 6s¹ rather than [Xe] 4f¹⁴ 5d⁹ 6s². These deviations maximize Hund's rule by promoting parallel spins in half-filled subshells or achieving complete d-subshell filling, which lowers overall energy through reduced electron-electron repulsion and increased exchange interactions. In heavier transition metals like , relativistic effects further contribute by contracting the orbital and expanding the 5d orbitals, making the 6s-5d energy separation even smaller and favoring the d¹⁰ . This pattern of exceptions is predictable, occurring primarily at groups 6 (, ), 7 (, ), 10 (), 11 (, , ), and 12 boundaries where half- or full-filling aligns with valence electron counts. Such anomalies affect approximately 10% of the 40 d-block elements ( to , Y to Cd, La-Hf to ), highlighting the principle's limitations in multi-electron systems where subshell interactions dominate. For chromium, the [Ar] 4s¹ 3d⁵ configuration is evidenced by its first ionization energy of 652.9 kJ/mol, which is lower than expected for removing an electron from a paired 4s² orbital, as the process yields the stable Cr⁺ ([Ar] 3d⁵) half-filled subshell. Similarly, copper's first ionization energy of 745.5 kJ/mol is notably lower than zinc's 906.4 kJ/mol, reflecting the ease of removing the single 4s electron to form Cu⁺ ([Ar] 3d¹⁰), a fully filled subshell that resists further ionization (second IE: 1957.9 kJ/mol). These ionization patterns confirm the ground-state configurations spectroscopically and energetically.

Anomalies in Lanthanides and Actinides (f-block)

In the f-block elements, comprising the lanthanides (elements 57–71) and actinides (89–103), the Aufbau principle anticipates sequential filling of the 4f orbitals for lanthanides after the 5d and preceding the 6s, but numerous deviations arise due to close energy proximities among the 4f, 5d, and 6s subshells. For instance, lanthanum (La) adopts the configuration [Xe] 5d¹ 6s² rather than the expected [Xe] 6s², as the 5d orbital is energetically favorable. Cerium (Ce) exhibits [Xe] 4f¹ 5d¹ 6s² in its ground state, though a low-lying excited state [Xe] 4f² 6s² exists, reflecting competition between 4f and 5d filling. Gadolinium (Gd) follows with [Xe] 4f⁷ 5d¹ 6s² instead of [Xe] 4f⁸ 6s², prioritizing the half-filled 4f subshell for stability. Lutetium (Lu) concludes the series with [Xe] 4f¹⁴ 5d¹ 6s² over [Xe] 4f¹⁴ 6s², again favoring 5d involvement. Similar irregularities occur in the actinides, where relativistic effects further complicate orbital energies due to high nuclear charge. (Ac) has [Rn] 6d¹ 7s², bypassing 5f filling, while (Th) shows [Rn] 6d² 7s² rather than [Rn] 5f⁰ 7s². These deviations stem primarily from the poor shielding efficacy of f electrons, which are radially extended and fail to effectively screen the nuclear charge from outer electrons, resulting in a higher that stabilizes 5d and 6s (or 6d and 7s) orbitals over 4f (or 5f). In actinides, scalar relativistic effects contract the 6s and 5d orbitals while expanding 7p, exacerbating the energy overlap and promoting d over f filling in early members. The lanthanide contraction exemplifies these shielding issues, wherein atomic and ionic radii decrease progressively across the series (from 1.03 Å for La³⁺ to 0.86 Å for Lu³⁺, Shannon effective ionic radii, 6) more sharply than in d-block trends, as the added electrons provide minimal shielding against the increasing nuclear charge. This contraction influences orbital filling by raising energies relative to 5d in later elements and impacts post-lanthanide properties, such as smaller radii in sixth-period transition metals. For elements like (Pr) and (Nd), ground states are [Xe] 4f³ 6s² and [Xe] 4f⁴ 6s², respectively, though low-lying excited states exist (e.g., [Xe] 4f² 5d¹ 6s² for Pr); these are confirmed through spectroscopic of levels. Despite these anomalies, the +3 oxidation state dominates across the lanthanides, achieved by loss of the 6s² electrons and one from 5d or 4f, yielding stable [Xe] 4fⁿ cores that resist further ionization due to the contracted, shielded 4f subshell. In actinides, relativistic stabilization of 7s² similarly favors +3 for early members like Ac, though higher states emerge later.

Historical Context

Emergence in Early Quantum Mechanics

The foundations of the Aufbau principle emerged from early attempts to model atomic structure in the context of nascent , beginning with Bohr's 1913 model of the . In this framework, electrons occupy discrete, stationary orbits around the nucleus, with quantized preventing classical and explaining atomic spectral lines. Bohr's model initially addressed single-electron systems but laid the groundwork for conceptualizing electron arrangements in multi-electron atoms by suggesting that additional electrons would fill successive orbits of increasing energy. Arnold Sommerfeld extended Bohr's circular orbits in 1916 by introducing elliptical orbits to account for relativistic effects and in lines, incorporating a second quantum number to describe . This refinement allowed for a more nuanced description of energy levels in hydrogen-like atoms and provided a basis for extending the model to multi-electron systems, where electrons could occupy varied orbital shapes within discrete shells. Prior to these quantum developments, in the , Johannes Rydberg had correlated periodic table patterns with series, proposing that atomic properties arose from layered electron arrangements corresponding to distinct energy groups observed in emission spectra. The building-up process central to the Aufbau principle took shape through the independent proposals of Charles R. Bury in 1921 and Edmund C. Stoner in 1924, who suggested that electrons fill shells sequentially by order of increasing , drawing on Sommerfeld's s to predict configurations. Bury similarly advocated for layered shells with capacity limits, reinforcing the idea of progressive filling to explain and . Stoner's formulation emphasized that the maximum electrons per subshell relate to the inner , enabling a systematic distribution that aligned with observed chemical periodicity. This approach remained empirical, particularly for lighter elements, as precise orderings were not yet derived theoretically. A pivotal advancement came in with Wolfgang Pauli's introduction of the exclusion principle, which stipulated that no two electrons in an atom can share identical quantum numbers, thereby enforcing stable, non-degenerate shell filling and resolving inconsistencies in spectral intensities and stability. Pauli's rule complemented the building-up concept by limiting occupancy, but early models still lacked a rigorous criterion for orbital energy sequencing, such as the later n + l rule, rendering predictions approximate for complex atoms.

Refinements and Spectroscopic Evidence

The formalization of the Aufbau principle advanced significantly with the development of wave mechanics in the mid-1920s. In 1926, introduced his eponymous equation, which provided exact solutions for the , yielding wavefunctions that defined atomic orbitals characterized by quantum numbers n, l, and m_l. These orbitals formed the basis for understanding arrangements in multi-electron atoms, as the equation's separable form in spherical coordinates allowed for the identification of energy levels dependent primarily on n for hydrogen-like systems. To extend this to multi-electron atoms, Douglas R. Hartree developed the self-consistent field (SCF) method in 1928, approximating the by treating each as moving in an average potential created by the and other electrons. This iterative approach solved the numerically for atoms like sodium and , producing orbital energies that aligned with the building-up order observed in spectra, though it initially neglected electron exchange effects. Erwin Madelung formalized the n + l rule in 1936 within the framework of SCF theory, deriving an energy ordering where subshell energies increase with the sum n + l, and for equal sums, with n. This empirical yet theoretically grounded rule explained the sequence of orbital filling as arising from the balance of kinetic and potential energies in self-consistent atomic fields, providing a predictive tool for electron configurations beyond . Spectroscopic evidence for this ordering came from emission lines and optical spectra analyzed by and collaborators in the 1920s and 1930s, which confirmed the Aufbau sequence up to (Z=36) through observed term energies and ionization potentials matching the predicted shell closures, such as the filling of 3p orbitals in argon-like cores. Key refinements included Friedrich Hund's 1927 work on atomic multiplet structure, which incorporated electron and Pauli exclusion to determine ground-state terms, ensuring maximum spin multiplicity for degenerate orbitals and supporting the stability of half-filled subshells in the Aufbau process. John C. Slater's 1930 introduction of orthogonalized atomic orbitals addressed limitations in SCF approximations by enforcing between core and valence functions, helping explain deviations from strict n + l ordering through effective shielding and effects. In the modern era, (DFT), building on Hohenberg-Kohn theorems since 1964, refines these ideas by optimizing total rather than orbitals directly, yet computational DFT calculations consistently uphold the Aufbau principle as a reliable for ground-state configurations across the periodic table, with exceptions arising from effects quantifiable via .

Applications and Implications

Building the Periodic Table

The Aufbau principle provides the foundational framework for constructing the modern periodic table by dictating the sequential filling of atomic orbitals with electrons as increases, thereby organizing elements into distinct blocks based on the subshell involved in their valence configurations. The s-block comprises elements with valence electrons in the ns¹ or ns² orbitals, typically the and alkaline earth metals in groups 1 and 2. The p-block includes elements with configurations ns² np¹⁻⁶, encompassing groups 13 through 18 and featuring a wide range of nonmetals, metalloids, and post-transition metals. The d-block, or transition metals, arises from ns² (n-1)d¹⁻¹⁰ configurations, spanning groups 3 through 12. Finally, the f-block consists of the lanthanides and actinides with ns² (n-2)f¹⁻¹⁴ (n-1)d⁰⁻¹ configurations, where the f subshell filling dominates the inner layers. This orbital filling sequence directly determines the lengths of the periodic table's periods, reflecting the maximum electron capacities of the subshells: the first period holds 2 elements (1s²), the second and third periods each accommodate 8 elements (2s²2p⁶ and 3s²3p⁶, respectively), the fourth and fifth periods expand to 18 elements each (incorporating 4s²3d¹⁰4p⁶ and 5s²4d¹⁰5p⁶), and the sixth and seventh periods reach 32 elements due to the insertion of the 14-electron f subshells (e.g., 4f¹⁴ in the sixth period alongside 6s²5d¹⁰6p⁶). These capacities—2 for s, 6 for p, 10 for d, and 14 for f—arise from the combined with Aufbau ordering, ensuring the table's rows align with n while accommodating the delayed filling of d and f orbitals. Group assignments in the periodic table stem from the number of valence electrons in the outermost s and p orbitals, as per the Aufbau principle, which governs chemical similarity within columns. For the main-group elements (groups IA through VIIIA, or 1, 2, and 13–18), the group number corresponds directly to the total s + p valence electrons (e.g., group IA has ns¹, group VIIIA has ns² np⁶). In the transition metal groups (IB through VIIIB, or 3–12), the outer s electrons and partial d involvement influence bonding, but the primary valence count from s/p still underpins the grouping, though with nuances due to variable oxidation states. Exceptions to strict Aufbau filling, such as in the d-block where elements like ([Ar] 4s¹ 3d⁵) and ([Ar] 4s¹ 3d¹⁰) prioritize half-filled or fully filled subshells for stability, introduce irregularities that subtly affect properties like and but do not alter the overall block structure. In the f-block, the insertion of the 14 and elements as separate series below the main table accommodates the poor shielding by f electrons, preventing excessive lengthening of periods 6 and 7 while highlighting their distinct inner-transition characteristics. This layout traces back to Mendeleev's periodic table, whose empirical groupings by atomic weight and properties were later rationally explained by filling under the Aufbau principle, bridging classical and quantum understandings of periodicity.

Predicting Atomic and Molecular Properties

The Aufbau principle governs the sequential filling of atomic orbitals as the atomic number (Z) increases, which directly influences periodic trends in atomic properties such as radii and ionization energies. As electrons occupy successively higher energy levels, atomic radii generally increase down a group due to the addition of new principal quantum shells, while they decrease across a period owing to the increasing effective nuclear charge (Z_eff) that pulls electrons closer to the nucleus. Ionization energy (IE), the energy required to remove an electron, follows similar patterns: it decreases down a group as valence electrons are shielded by inner shells and reside farther from the nucleus, but increases across a period due to higher Z_eff and stable subshell configurations. Representative examples illustrate these trends rooted in Aufbau-derived configurations. Noble gases, with fully filled p⁶ subshells (e.g., Ne: [He] 2s²2p⁶), exhibit exceptionally high first IE values, such as 2081 kJ/mol for Ne, due to the stability of their closed-shell structures that resist electron removal. In contrast, alkali metals with ns¹ valence configurations (e.g., Na: [Ne] 3s¹) have low IE, like 496 kJ/mol for Na, facilitating easy loss of the outer electron to form positive ions and enabling metallic bonding through delocalized electrons. These properties underscore how orbital filling dictates chemical reactivity and bonding tendencies. At the molecular level, the valence electron count from Aufbau configurations determines bonding patterns, including Lewis structures, VSEPR theory, and hybridization. Valence electrons, typically those in the outermost s and p orbitals, dictate the number of bonds an atom forms to achieve an octet; for instance, carbon's [He] 2s²2p² configuration leads to four bonds in methane (CH₄) via sp³ hybridization, predicting a tetrahedral geometry under VSEPR. In transition metals, partially filled d orbitals contribute additional valence electrons, enabling diverse coordination and reactivity, such as in catalysis where d electrons facilitate substrate activation and electron transfer in processes like olefin polymerization. Anomalies in d-block electron configurations, arising from Aufbau exceptions, further explain variable oxidation states that enhance molecular versatility. For example, chromium's configuration ([Ar] 4s¹ 3d⁵) allows stable half-filled d subshells, supporting oxidation states like +3 (losing 4s and three 3d electrons) and +6 (losing all electrons), which manifest in compounds such as Cr₂O₃ (antiferromagnetic) and chromates (oxidizing agents). These states arise because d electrons can be selectively removed or involved in without disrupting . In , the Aufbau principle provides an initial guess for orbital occupations in self-consistent field (SCF) calculations, accelerating convergence to ground-state molecular orbitals. By assigning electrons to lowest-energy molecular orbitals per Aufbau, methods like Hartree-Fock or initialize the , enabling efficient prediction of molecular properties such as energies and geometries, though adjustments account for non-Aufbau solutions in excited states.

References

  1. [1]
    Aufbau Principle - Chemistry 301
    We need to keep in mind that two electrons can go into each orbital. Thus a 3d subshell which has 5 d orbitals can "hold" 10 electrons.
  2. [2]
    Quantum Numbers and Electron Configurations
    The basis of this prediction is a rule known as the aufbau principle, which assumes that electrons are added to an atom, one at a time, starting with the lowest ...
  3. [3]
    Electron Configurations - FSU Chemistry & Biochemistry
    ... orbitals is based on the order of their energy. This is referred to as the Aufbau principle. The lowest energy orbitals fill first. Just like the quantum ...
  4. [4]
    30.9 The Pauli Exclusion Principle – College Physics chapters 1-17
    The Pauli exclusion principle explains why some configurations of electrons are allowed while others are not. Since electrons cannot have the same set of ...
  5. [5]
    Hund's Rule - Chemistry 301
    Hund's Rule. The Aufbau principle let's us build up an atoms electronic configuration by placing electrons into orbitals of every increasing energy.Missing: origin | Show results with:origin
  6. [6]
    Orbitals and Quantum Numbers | CK-12 Foundation
    The principal quantum number (n) defines the energy level of an electron. The azimuthal quantum number (l) determines the orbital shape. The magnetic quantum ...
  7. [7]
    Pauli Exclusion Principle - HyperPhysics
    No two electrons in an atom can have identical quantum numbers. This is an example of a general principle which applies not only to electrons but also to other ...
  8. [8]
    Many-Electron Atoms - The Electronic Basis of the Periodic Table
    For the many-electron atoms, the energy of an orbital depends on both n and l, the energy increasing as l increases even when n is constant. For example, from ...
  9. [9]
    s, p, d, f Atomic Orbitals - Chemistry Steps
    S orbitals have a spherical shape, p orbitals are dumbbell-shaped, d orbitals are shaped like a cloverleaf, and f orbitals are characterized by more complex ...
  10. [10]
    [PDF] ORDINAL EXPLANATION OF THE PERIODIC SYSTEM OF ...
    for two states with the same value of n+l, the one with larger n has the larger energy. This rule of Bohr's is also called the Aufbau rule, or the Madelung rule ...
  11. [11]
    [PDF] The angular momentum distribution of electrons in large atoms and ...
    Jun 17, 2025 · It is a well established rule in chemistry that the ordering of subshells in neutral atoms is given by the Madelung rule, also known as the.<|control11|><|separator|>
  12. [12]
    [PDF] Approximating Hamiltonian for Hartree-Fock solutions for ... - arXiv
    Jun 17, 2025 · It can be derived by applying the quasi-classical approximation in solving the Hartree-Fock equations for a nonrelativistic multi- electron ...
  13. [13]
    The Löwdin challenge: Origin of the n+ℓ, n (Madelung) rule for filling ...
    Jun 10, 2002 · The n+ℓ , n (Madelung) rule is derived from the perturbation operator ... * Herman and Skillman's Hartree–Fock–Slater atomic structure ...
  14. [14]
  15. [15]
    Electronic Orbitals - Chemistry LibreTexts
    Jan 29, 2023 · This means that the s orbital can contain up to two electrons, the p orbital can contain up to six electrons, the d orbital can contain up to 10 ...
  16. [16]
    Periodic Table Blocks of Elements - Science Notes
    Sep 21, 2020 · Periodic table blocks are sets of elements grouped by their valence electron orbitals. The four block names are s-block, p-block, d-block, and f-block.
  17. [17]
  18. [18]
    Atoms: Aufbau Process and Electron Filling
    ### Summary of Reasons for Exceptions in Transition Metal Electron Configurations
  19. [19]
    The Challenge of the So‐Called Electron Configurations of the ...
    May 9, 2006 · Chemically bound atoms differ from free atoms significantly: Diffuse Rydberg orbitals, for example, (n+1)s, are perturbed away in compounds.
  20. [20]
    first ionisation energy - Chemguide
    All of these elements have an electronic structure [Ar]3dn4s2 (or 4s1 in the cases of chromium and copper). The electron being lost always comes from the 4s ...
  21. [21]
    Atomic Data for Lanthanum (La) - Physical Measurement Laboratory
    Atomic Data for Lanthanum (La) Atomic Number = 57 Atomic Weight = 138.9055 Reference E95 La I Ground State (1s 2 2s 2 2p 6 3s 2 3p 6 3d 1Missing: configurations | Show results with:configurations
  22. [22]
    [PDF] 5. Electronic Structure of the Elements - Particle Data Group
    May 31, 2024 · The electronic configurations, ground state levels, and ionization ... 58 Ce Cerium. (Xe)4f 5d 6s2. 1G◦. 4. 5.5386. 59 Pr Praseodymium(Xe)4f3.
  23. [23]
    Atomic Data for Gadolinium (Gd) - Physical Measurement Laboratory
    NIST Physical Measurement Laboratory, Handbook of Basic Atomic Spectroscopic ... Gd I Ground State (1s22s22p63s23p63d104s24p64d105s25p6)4f75d6s2 9D°2 ...
  24. [24]
    Atomic Data for Lutetium (Lu) - Physical Measurement Laboratory
    Atomic Data for Lutetium (Lu) Atomic Number = 71 Atomic Weight = 174.967 Reference E95 Lu I Ground State (1s 2 2s 2 2p 6 3s 2 3p 6 3d 1Missing: electron configuration
  25. [25]
    Atomic Data for Actinium (Ac) - Physical Measurement Laboratory
    Ac I Ground State 1s22s22p63s23p63d104s24p64d104f145s25p65d106s26p66d7s2 2D3/2. Ionization energy 41700 cm-1 (5.17 eV) Ref. S74 Ac II Ground State ...Missing: electron configuration thorium
  26. [26]
    Atomic Data for Thorium (Th) - Physical Measurement Laboratory
    Atomic Data for Thorium (Th) Atomic Number = 90 Atomic Weight = 232.0381 Reference E95 Th I Ground State 1s 2 2s 2 2p 6 3s 2 3p 6 3d 1Missing: electron configuration actinium Ac
  27. [27]
    [PDF] Actinide Research Quarterly - Los Alamos National Laboratory
    The primary f physical cause of the actinide (and lanthanide) contraction is decreased shielding of the core positive charge across the series: as electrons.
  28. [28]
    Relativistic Effects in the Electronic Structure of Atoms | ACS Omega
    Sep 22, 2017 · Periodic trends in relativistic effects are investigated from 1 H through 103 Lr using Dirac–Hartree–Fock and nonrelativistic Hartree–Fock calculations.Introduction · Results and Discussion · Supporting Information · References
  29. [29]
    [PDF] Systematic study of the phase behavior of f-block oxides irradiated ...
    The poor screening provided by f-electrons results in a monotonic decrease in the ionic radius of Ln3+ across the series, a behavior known as the lanthanide ...
  30. [30]
    [PDF] Atomic energy levels - the rare-earth elements
    This document covers atomic energy levels for rare-earth elements from Lanthanum (Z=57) through Lutetium (Z=71), including Cerium, Praseodymium, Neodymium, ...
  31. [31]
    Niels Bohr – Facts - NobelPrize.org
    In 1913, Niels Bohr proposed a theory for the hydrogen atom, based on quantum theory that some physical quantities only take discrete values. Electrons move ...Missing: multi- | Show results with:multi-
  32. [32]
    2.3: Atomic Orbitals and the Bohr Model - Chemistry LibreTexts
    Aug 1, 2020 · The Bohr model of the hydrogen atom explains the connection between the quantization of photons and the quantized emission from atoms.
  33. [33]
    Sommerfeld's elliptical atomic orbits revisited—A useful preliminary ...
    Sep 1, 2014 · Sommerfeld's extension of the Bohr Model1 to elliptical orbits in 1915, later described in detail in his book Atomic Structure and Spectral ...Missing: multi- | Show results with:multi-
  34. [34]
    Getting the numbers right - the lonely struggle of Rydberg | Feature
    From this model emerged the idea of concentric electron 'shells', mirroring the recurring patterns in the periodic table of the elements. Electron jumps between ...Missing: 1910s | Show results with:1910s
  35. [35]
    Full article: LXXIII. The distribution of electrons among atomic levels
    Apr 8, 2009 · A distribution of electrons in the atom is proposed, according to which the number in a sub-group is simply related to the inner quantum number characterizing ...Missing: shell | Show results with:shell
  36. [36]
    Charles Bury on Electronic Structure in 1921 - chemteam.info
    The maximum number of electrons in each shell or layer is proportional to the area of its surface; thus, successive layers can contain 2, 8, 18 and 32 electrons ...Missing: 1924 energy
  37. [37]
    [PDF] On the Connexion between the Completion of Electron Groups in an ...
    Apr 16, 2010 · PAULI. Z. Physik 31, 765ff (1925). Especially in connexion with Millikan and Landé's observation that the alkali doublet can be represented ...
  38. [38]
    January 1925: Wolfgang Pauli announces the exclusion principle
    Pauli originally applied the exclusion principle to explain electrons in atoms, but later it was extended to any system of fermions, which have half integer ...Missing: URL | Show results with:URL
  39. [39]
    [PDF] 1926-Schrodinger.pdf
    At first sight this equation seems to offer ill means of solving atomic problems, e.g. of defining discrete energy-levels in the hydrogen atom. Being a partial ...
  40. [40]
    Density functional theory: Its origins, rise to prominence, and future
    Aug 25, 2015 · This paper reviews the development of density-related methods back to the early years of quantum mechanics and follows the breakthrough in their application ...
  41. [41]
    Periodic Trends - EdTech Books
    The first ionization energy of the elements in the first five periods are plotted against their atomic number. This version of the periodic table shows the ...
  42. [42]
    [PDF] LECTURE 5. PERIODIC TRENDS EXPLAINED BY EFFECTIVE ...
    Rule 1: Effective nuclear charge (ENC) will explain the relative size and interest in electrons for atoms and ions. As will be shown, for example, as ENC⇩ Size ...
  43. [43]
    [PDF] The periodic system of the elements Predict (rather than passively ...
    ➢. As a result of Hund's rule, constraints are placed on the way atomic orbitals are filled using the aufbau principle. Before any two electrons occupy an ...<|control11|><|separator|>
  44. [44]
    Electron Configurations & The Periodic Table - MSU chemistry
    According to the Aufbau principle, the electrons of an atom occupy quantum levels or orbitals starting from the lowest energy level, and proceeding to the ...
  45. [45]
    Catalytically Influential Features in Transition Metal Oxides
    Nov 2, 2021 · In TMOs, the d-electrons of transition metals play the major role in shaping the electronic structure of TMOs. The ...
  46. [46]
    Oxidation States of Transition Metals - Chemistry LibreTexts
    Jun 30, 2023 · For example, in group 6, (chromium) Cr is most stable at a +3 oxidation state, meaning that you will not find many stable forms of Cr in the +4 ...
  47. [47]
    [PDF] A Simple and Robust Approach to Orbital Relaxation of Non-Aufbau ...
    Jul 9, 2020 · The Aufbau principle is a basic tenet of ground-state electronic structure theory, but many chemical properties such as ionization energies and ...Missing: origin | Show results with:origin