Fact-checked by Grok 2 weeks ago

Electron shell

In and chemistry, an electron shell refers to a group of atomic orbitals that share the same , n, which defines the and average distance of electrons from the in an atom. These shells organize the electrons surrounding the positively charged , with electrons filling lower-energy shells first according to the . The concept originated in the of the atom but is now understood through , where shells correspond to discrete energy states rather than fixed circular orbits. Electron shells are traditionally labeled with letters starting from the innermost: K (n=1), L (n=2), M (n=3), and so on, up to the seventh shell for known elements. Each shell can hold a maximum of 2n² electrons, allowing the first shell to accommodate 2 electrons, the second up to 8, the third up to 18, and higher shells even more, though outer shells in most elements are not fully occupied. This capacity arises from the subshells (s, p, d, f) within each shell, each with specific orbital quantum numbers that determine their electron-holding limits. The arrangement of electrons in shells governs an atom's chemical behavior, particularly through the valence shell—the outermost shell—which contains the valence electrons responsible for bonding and reactivity. Full valence shells, as in , confer stability, while incomplete ones drive elements to form ions or compounds to achieve a stable octet configuration in the valence shell. This shell structure underpins the periodic table's organization, where periods correspond to the filling of successive shells, explaining trends in atomic size, , and across elements.

Fundamentals

Definition and Basic Properties

In atomic physics, electron shells represent discrete energy levels within an atom, conceptualized as concentric spherical regions surrounding the nucleus where electrons are most likely to be found according to quantum mechanical probability distributions. These shells are defined by the principal quantum number n, a positive integer that specifies the shell's designation, with n = 1 for the innermost shell and increasing sequentially outward as n = 2, 3, 4, \dots. The value of n indicates the shell's average distance from the nucleus and its associated energy, with higher n corresponding to larger, higher-energy shells. Electron shells are conventionally labeled using : the K shell for n=1, L shell for n=2, M shell for n=3, N for n=4, and so forth. A key property is that shell radius and binding both increase with n, meaning in outer shells are less tightly bound to the and require less energy to remove. Consequently, the occupying the outermost shell—termed valence electrons—predominantly govern an atom's chemical behavior, including its reactivity and bonding tendencies with other atoms. For conceptual clarity, electron shells can be intuitively compared to the discrete orbits in a simplified planetary model of the atom, akin to planets revolving around the sun at fixed distances; this analogy, however, oversimplifies the quantum reality, where electrons do not follow definite paths but exist in probabilistic clouds. Each shell is subdivided into subshells based on the .

Role in Atomic and Molecular Structure

Electron shells play a crucial role in determining the stability of atoms by organizing electrons into energy levels that minimize overall energy. Atoms with completely filled inner shells, such as the noble gases, exhibit exceptional stability due to their closed-shell configurations, where the valence shell is fully occupied, leading to low reactivity and high ionization energies. For instance, helium achieves this stability with its two electrons filling the 1s shell completely, resulting in a configuration that resists chemical interactions under normal conditions. In chemical bonding, electron shells govern how atoms interact to achieve more stable , primarily through the sharing or transfer of from the outermost shell. The describes the tendency of atoms to gain, lose, or share electrons to fill their outer shell with eight electrons, mimicking the stable arrangement and lowering . This leads to , where electrons are transferred—as in sodium, which loses its single 3s to form Na⁺ with a neon-like —creating oppositely charged ions attracted by electrostatic forces. In covalent bonding, atoms share pairs of to satisfy the , forming molecules where the shared electrons are counted toward both atoms' outer shells. The arrangement of electrons in outer shells also influences molecular structure by dictating bond angles and geometries through interactions of valence electrons. In molecular formation, these valence electrons participate in hybridization, where atomic orbitals from the outer shell combine to form hybrid orbitals that overlap effectively, enabling specific molecular shapes like the tetrahedral geometry in . This shell-driven hybridization and repulsion among electron pairs in the valence shell determine overall molecular architecture, affecting properties such as and reactivity.

Historical Context

Pre-Quantum Models

In the early , J.J. Thomson proposed the of the atom in 1904, envisioning it as a uniform sphere of positive charge with negatively charged embedded throughout like plums in a pudding, ensuring overall electrical neutrality without any discrete shell structure. This model accounted for the discovery of the but lacked a concept of organized electron layers, treating electrons as distributed within a diffuse positive medium. Ernest Rutherford's nuclear model, introduced in 1911 based on the gold foil scattering experiments conducted by and , depicted the as a tiny, dense, positively charged surrounded by electrons orbiting in planetary-like paths, much like planets around the sun. However, this classical picture was inherently unstable, as accelerating electrons in circular orbits would, according to Maxwell's electromagnetic theory, continuously radiate energy and spiral inward toward the , leading to atomic collapse. Niels Bohr addressed these issues in his 1913 model, postulating that occupy discrete, quantized orbits around the nucleus—serving as precursors to modern electron shells—where they do not radiate while in these stable "stationary states." Bohr introduced quantization, with the electron's orbital given by L = n \hbar, where n is a positive and \hbar = h / 2\pi (with h as Planck's constant), preventing continuous loss. This framework successfully explained the discrete spectral lines of , attributing them to electrons transitioning between quantized levels and emitting photons of specific wavelengths corresponding to the energy differences. Despite its successes, Bohr's model had significant limitations, particularly its inability to accurately describe multi-electron atoms, where electron-electron interactions were not accounted for, leading to incorrect predictions of energy levels and spectra for elements beyond . This shortfall highlighted the need for a more comprehensive quantum mechanical treatment to incorporate wave-like behavior and full atomic complexity.

Development of Quantum Theory

The development of marked a pivotal shift in understanding electron behavior within atoms, building upon the limitations of earlier atomic models such as Niels Bohr's 1913 planetary model, which posited quantized electron orbits but failed to fully explain spectral complexities. A key milestone in this evolution came from in 1916, who extended Bohr's model by incorporating relativistic effects and allowing for elliptical orbits rather than strictly circular ones. This introduction of an additional quantum condition for the radial motion, alongside angular quantization, provided a precursor to the concept of subshells by accounting for splittings in atomic spectra, such as the hydrogen fine structure lines observed experimentally. Sommerfeld's relativistic treatment predicted energy level shifts dependent on orbital eccentricity, laying groundwork for more nuanced electron groupings in multi-electron atoms. Further advancements arose from Louis de Broglie's 1924 hypothesis, which proposed wave-particle duality for electrons, asserting that particles like electrons exhibit wave-like properties with a wavelength given by \lambda = h / p, where h is Planck's constant and p is the electron's momentum. This idea suggested that electrons in atoms could be standing waves confined to discrete paths, providing a conceptual bridge from classical orbits to wave mechanics and motivating the quantization of electron shells. Building on this, Erwin Schrödinger formulated the time-independent Schrödinger equation in 1926 for the hydrogen atom, expressed as \hat{H} \psi = E \psi, where \hat{H} is the Hamiltonian operator, \psi is the wave function, and E is the energy eigenvalue. Solving this equation yielded quantized energy levels E_n = -13.6 \, \text{eV} / n^2, where n is a positive , confirming discrete shells and resolving inconsistencies in Bohr's non-relativistic energies. Complementing these developments, Werner Heisenberg's uncertainty principle, articulated in 1927, introduced fundamental indeterminacy in quantum measurements, stating that the product of uncertainties in position \Delta x and momentum \Delta p satisfies \Delta x \Delta p \geq \hbar / 2, where \hbar = h / 2\pi. This principle prohibited the precise definition of fixed electron orbits, as simultaneous knowledge of position and momentum becomes impossible, thereby necessitating a probabilistic wave description for electron shells rather than deterministic paths. Finally, Wolfgang Pauli's exclusion principle, proposed in 1925, stipulated that no two electrons in an atom can occupy the same quantum state, meaning they cannot share all quantum numbers simultaneously. This rule was essential for establishing the shell structure, as it enforced distinct occupancy limits within energy levels, preventing collapse of electrons into the lowest state and enabling the layered configuration observed in atomic spectra.

Quantum Mechanical Framework

Principal Quantum Number and Shells

In , the n is a (n = 1, 2, 3, \dots, \infty) that primarily determines the and size of an electron shell in an atom. It labels the discrete energy shells, with n = 1 corresponding to the innermost K shell, n = 2 to the L shell, n = 3 to the M shell, and so on, following established in early . The energy of an electron in a shell depends on n. For hydrogen-like atoms (single-electron systems such as H or He⁺), the energy levels are given by E_n = -\frac{13.6 \, Z^2}{n^2} \, \text{eV}, where Z is the atomic number; this formula arises from solving the Schrödinger equation for the Coulomb potential. In multi-electron atoms, inner electrons screen the nuclear charge, reducing the effective nuclear charge Z_{\text{eff}} experienced by outer electrons, which modifies the energy to approximately E_n \approx -\frac{13.6 \, Z_{\text{eff}}^2}{n^2} \, \text{eV} and results in less negative (higher) energies compared to hydrogen-like cases. Spatially, the electron shell's extent is characterized by the radial , where the probability for finding the peaks at an average distance from the scaling as \sim n^2 a_0, with a_0 = 0.529 \, \text{Å} being the (the most probable radius for the of ). This quadratic dependence on n reflects the increasing orbital size with higher shells, as derived from the radial wavefunctions in the solution. Electron shells are denoted using spectroscopic notation, such as 1s or 2p, where the numeral specifies n and the letter indicates the subshell (an angular subdivision within the shell); this ties the overall shell structure directly to the principal quantum number.

Subshells and Angular Momentum

Within each electron shell defined by the principal quantum number n, subshells arise as subdivisions characterized by the azimuthal quantum number l, which quantifies the orbital angular momentum of the electron. The value of l ranges from 0 to n-1 in integer steps, with this range constrained by n. Subshells are conventionally labeled using letters: s for l = 0, p for l = 1, d for l = 2, and f for l = 3, with higher values of l assigned subsequent letters in alphabetical order (g, h, etc.). The orbital associated with a subshell has a magnitude given by \sqrt{l(l+1)} \hbar, where \hbar = h / 2\pi and h is Planck's constant; this formula emerges from the quantum mechanical solution to the angular part of the for the . The distinct values of l also determine the shapes of the atomic orbitals within each subshell. For instance, s orbitals (l = 0) are spherically symmetric around the , p orbitals (l = 1) exhibit a shape with two lobes along a principal axis, d orbitals (l = 2) possess more intricate cloverleaf-like patterns with four lobes in the equatorial plane or double dumbbells, and f orbitals (l = 3) display even more complex geometries involving multiple lobes and nodal planes. These shapes reflect the probability distribution of finding the electron in space, derived from the angular dependence of the wave function. Further subdivision within a subshell is provided by the m_l, which specifies the orientation of the orbital in a and takes integer values from -l to +l, inclusive. This yields $2l + 1 possible values for m_l, corresponding to the number of distinct orbitals in the subshell; each orbital represents a unique spatial orientation. For example, the p subshell (l = 1) has three orbitals (m_l = -1, 0, +1), oriented along the x, y, and z axes, while the d subshell (l = 2) contains five orbitals.

Orbital Energies and Electron Filling

In the , the energy levels of orbitals depend solely on the n, resulting in degeneracy where all subshells within the same shell (e.g., 2s and 2p) have identical . However, in multi-electron , this degeneracy is lifted primarily by electron-electron repulsion, which introduces additional contributions that differentiate orbitals based on both n and the l. As a result, orbital increase with increasing n due to the larger average distance from the , but within a given n, also rise with l because higher-l orbitals experience less penetration toward the and thus weaker attraction. For example, the 2s orbital has lower than the 2p orbital in like carbon, as the 2s wavefunction penetrates closer to the , allowing better shielding from inner electrons and a stronger effective pull. This variation in orbital energies arises from the effective nuclear charge Z_{\text{eff}} experienced by an , defined as Z_{\text{eff}} = Z - \sigma, where [Z](/page/Z) is the (nuclear charge) and \sigma is the screening constant accounting for the of inner electrons. The screening constant \sigma quantifies how much the net positive charge is reduced by electron-electron repulsions and the spatial distribution of other electrons, with values estimated via methods like , which assign contributions based on the electron's shell and subshell grouping. Electrons in penetrating orbitals (lower l) experience a higher Z_{\text{eff}} because they are less effectively screened, lowering their energy relative to non-penetrating ones in the same shell. When filling orbitals with electrons, the limits each orbital to a maximum of two electrons, distinguished by their m_s, which can take values of +\frac{1}{2} (spin up) or -\frac{1}{2} (spin down). These opposite spins allow the two electrons to occupy the same orbital without violating the principle that no two electrons can have identical sets of quantum numbers. In degenerate orbitals (those of equal energy within a subshell), Hund's rule dictates that electrons occupy orbitals singly with parallel spins to maximize the total spin multiplicity, minimizing electron-electron repulsion through greater spatial separation before pairing occurs. This configuration achieves the lowest energy state by favoring higher total spin S, where the multiplicity $2S + 1 is maximized for the subshell.

Configuration Rules

Maximum Occupancy per Shell and Subshell

In , the maximum number of electrons that can occupy a subshell is determined by the \ell, given by the $2(2\ell + 1). This arises from the $2\ell + 1 possible values of the m_\ell (ranging from -\ell to +\ell), each defining a distinct orbital that can accommodate up to two electrons with opposite spins. For the common subshells, this yields specific limits: the s subshell (\ell = 0) holds 2 electrons, the p subshell (\ell = 1) holds 6 electrons, the d subshell (\ell = 2) holds 10 electrons, and the f subshell (\ell = 3) holds 14 electrons. These limits are enforced by the , which prohibits more than two electrons from sharing the same set of quantum numbers in an orbital. The total maximum occupancy of an electron shell, defined by the principal n, is $2n^2. This is derived by summing the capacities of all subshells within the shell, from \ell = 0 to \ell = n-1: \sum_{\ell=0}^{n-1} 2(2\ell + 1) = 2 \sum_{\ell=0}^{n-1} (2\ell + 1) = 2 \left[ 2 \cdot \frac{(n-1)n}{2} + n \right] = 2n^2. The summation reflects the total number of orbitals in the shell, which equals n^2, with each orbital holding 2 electrons. Examples of shell capacities include the K shell (n=1): 2 electrons; L shell (n=2): 8 electrons; and M shell (n=3): 18 electrons. In multi-electron atoms, shells generally fill from inner to outer due to decreasing electron-nucleus attraction with distance, though energy overlaps in transition metals can result in incomplete filling of inner d subshells before outer s subshells.
Principal Quantum Number (n)Shell DesignationMaximum Electrons ($2n^2)Subshells IncludedSubshell Capacities
1K21s2
2L82s, 2p2, 6
3M183s, 3p, 3d2, 6, 10
4N324s, 4p, 4d, 4f2, 6, 10, 14

Aufbau Principle and Madelung Rule

The Aufbau principle, also known as the building-up principle, dictates that electrons in the ground state of an atom occupy atomic orbitals in a sequence of increasing energy levels, starting from the lowest available energy orbital. This principle provides a systematic framework for constructing electron configurations by progressively adding electrons to orbitals as atomic number increases. To determine the specific order of orbital filling under the , the Madelung rule—also called the n + ℓ rule—is applied, where orbitals are filled in ascending order of the sum of the principal quantum number n and the ℓ (n + ℓ), and for orbitals with equal n + ℓ values, the one with the lower n is filled first. For example, the 1s orbital (n=1, ℓ=0; n+ℓ=1) is filled before the 2s orbital (n=2, ℓ=0; n+ℓ=2), which precedes the 2p orbitals (n=2, ℓ=1; n+ℓ=3), followed by 3s (n=3, ℓ=0; n+ℓ=3), 3p (n=3, ℓ=1; n+ℓ=4), 4s (n=4, ℓ=0; n+ℓ=4), and then 3d (n=3, ℓ=2; n+ℓ=5). While the Madelung rule accurately predicts the filling order for most elements, notable exceptions occur due to the enhanced stability of half-filled or fully filled subshells, which lower the overall energy compared to the predicted configuration. For instance, chromium (atomic number 24) adopts the configuration [Ar] 3d⁵ 4s¹ instead of the expected [Ar] 3d⁴ 4s², achieving a half-filled 3d subshell; similarly, copper (atomic number 29) has [Ar] 3d¹⁰ 4s¹ rather than [Ar] 3d⁹ 4s², resulting in a fully filled 3d subshell. In heavier elements following the lanthanide series, the lanthanide contraction—caused by the poor shielding of 4f electrons—further complicates the filling order by raising the energy of 5d and 4f orbitals relative to 6s, leading to irregularities in configurations for elements such as Hf and beyond. This diagonal arrangement, often depicted with arrows indicating the Aufbau path (1s → 2s → 2p → 3s → 3p → 4s → 3d → 4p), reflects the Madelung ordering and holds for most configurations up to , with the noted exceptions for .

Periodic Implications

Electron Distribution Across Elements

The periodic table is divided into four main blocks—s, p, d, and f—each corresponding to the subshell being filled by electrons in the configuration of atoms. The s-block includes groups 1 and 2 ( and metals), where the ns subshell accommodates up to two electrons. The p-block encompasses groups 13 through 18 (main group elements), filling the np subshell with up to six electrons. The d-block, spanning groups 3 through 12 ( metals), involves sequential filling of the (n-1)d subshell, which holds up to 10 electrons. The f-block consists of the lanthanides and actinides, where the (n-2)f subshell is filled, accommodating up to 14 electrons. This block organization reflects the progressive addition of electrons to specific subshells as increases, providing a structural basis for periodic properties. Shell filling trends vary across the blocks, influencing chemical behavior. In main group elements of the s- and p-blocks, electrons fill the outermost sequentially, leading to predictable valence electron counts that dictate reactivity. Transition metals in the d-block exhibit more complex filling, where the (n-1)d subshell is populated after the ns subshell but before completing the np subshell in subsequent periods, often resulting in variable oxidation states due to partial d-subshell occupancy. The f-block elements show similar intricacies, with f-subshell filling occurring within the inner shells, contributing to the contraction of atomic radii in these series. These trends are guided by the , which orders orbital filling by increasing energy. The valence shell, comprising the outermost occupied shell and any partially filled subshell, determines an element's group number and thus its position in the periodic table. For representative elements, the number of valence electrons corresponds directly to the group number; for example, group 17 elements () possess seven s in the configuration ns²np⁵. This valence electron count governs bonding tendencies, with s- and p-block elements typically achieving octet stability through or sharing. Ionization processes primarily affect the outer electron shells, as valence electrons are removed first to form cations, reflecting the relative stability of inner shells. Inner shells exhibit greater stability due to stronger nuclear attraction and shielding effects, making subsequent ionizations progressively more energy-intensive. Across the periodic table, first ionization energy increases from left to right within a period as effective nuclear charge rises and outer shells fill, while it decreases down a group with the addition of new shells that increase electron-nucleus distance. These shell-related effects underscore the periodic trends in reactivity and metallic character.

Notation and Examples for Key Elements

Electron configurations are expressed in spectroscopic notation, where the principal quantum number n precedes the subshell designation (s for l=0, p for l=1, d for l=2, f for l=3), followed by a superscript indicating the number of electrons in that subshell. For elements beyond helium, configurations often use noble gas core notation, enclosing the symbol of the preceding noble gas in brackets to represent its filled shells; for instance, neon's full configuration is $1s^2 2s^2 2p^6, while sodium's is abbreviated as [Ne] $3s^1. Representative examples demonstrate the progression of shell filling. Hydrogen has the simplest configuration: $1s^1. Carbon, in the second period, fills the 2p subshell partially: $1s^2 2s^2 2p^2. Iron, , involves d-orbital filling: $1s^2 2s^2 2p^6 3s^2 3p^6 4s^2 3d^6 or [Ar] $4s^2 3d^6. exhibits a complex configuration with f- and d-block involvement: [Xe] $4f^{14} 5d^{10} 6s^1. Exceptions to expected filling orders arise for enhanced stability. Chromium adopts [Ar] $4s^1 3d^5 rather than [Ar] $4s^2 3d^4, as the half-filled 3d subshell lowers energy through increased electron exchange interactions. Copper follows suit with [Ar] $4s^1 3d^{10} instead of [Ar] $4s^2 3d^9, prioritizing the fully filled 3d subshell for similar stability gains. The table below lists configurations for elements 1–20, revealing the sequential filling of K (n=1), L (n=2), M (n=3), and initial N (n=4) shells, with s subshells filling before p and the onset of 4s before 3d.
Atomic NumberElementElectron Configuration
1H$1s^1
2He$1s^2
3Li$1s^2 2s^1
4Be$1s^2 2s^2
5B$1s^2 2s^2 2p^1
6C$1s^2 2s^2 2p^2
7N$1s^2 2s^2 2p^3
8O$1s^2 2s^2 2p^4
9F$1s^2 2s^2 2p^5
10Ne$1s^2 2s^2 2p^6
11Na$1s^2 2s^2 2p^6 3s^1
12Mg$1s^2 2s^2 2p^6 3s^2
13Al$1s^2 2s^2 2p^6 3s^2 3p^1
14Si$1s^2 2s^2 2p^6 3s^2 3p^2
15P$1s^2 2s^2 2p^6 3s^2 3p^3
16S$1s^2 2s^2 2p^6 3s^2 3p^4
17Cl$1s^2 2s^2 2p^6 3s^2 3p^5
18Ar$1s^2 2s^2 2p^6 3s^2 3p^6
19K$1s^2 2s^2 2p^6 3s^2 3p^6 4s^1
20Ca$1s^2 2s^2 2p^6 3s^2 3p^6 4s^2

References

  1. [1]
    Atomic Spectros. - Atomic States, Shells, Configs - NIST
    Those electrons having the same principal quantum number n belong to the shell for that number. Electrons having both the same n value and l value belong to a ...
  2. [2]
    Quantum Numbers and Electron Configurations
    The principal quantum number (n) describes the size of the orbital. Orbitals for which n = 2 are larger than those for which n = 1, for example. Because they ...Quantum Numbers · Possible Combinations of... · Electron Configurations, the...
  3. [3]
    Background: Atoms and Light Energy - Imagine the Universe! - NASA
    Apr 20, 2020 · Surrounding the nucleus of an atom are shells of electrons - small negatively charged particles.
  4. [4]
    Crystal Chemistry - Tulane University
    Sep 22, 2014 · Electrons orbit around the nucleus in different shells, labeled from the innermost shell as K, L, M, N, etc. Each shell can have a certain ...
  5. [5]
    [PDF] Chapter 3. Many-Electron Atoms - Oregon State University
    ... Maximum number of electrons in a shell = 2n2 (closed or filled shell). Electrons having the same value of n and l are said to belong to the same subshell. - ...
  6. [6]
    Electron Configurations - FSU Chemistry & Biochemistry
    Electron configurations are the summary of where the electrons are around a nucleus. As we learned earlier, each neutral atom has a number of electrons equal ...
  7. [7]
    DOE Explains...Electrons - Department of Energy
    Atoms have multiple shells, each with a different number of subshells. Shells can hold more and more electrons the further they are from the nucleus, and ...
  8. [8]
    CH150: Chapter 2 - Atoms and Periodic Table - Chemistry
    This means that it has three electron shells that can house electrons. As electrons fill their available orbital spaces, they always fill the shells ...
  9. [9]
    Quantum Numbers and Electronic Structure
    The principal quantum number, n. The electronic energy levels in an atom are arranged roughly into principal levels (or shells) as specified by n. ... An electron ...
  10. [10]
    6.5: Quantum Numbers, Atomic Orbitals, and Electron Configurations
    The principal quantum number, signified by n, is the main energy level occupied by the electron. Energy levels are fixed distances from the nucleus of a given ...
  11. [11]
    The quantum mechanical view of the atom - Physics
    Aug 10, 1999 · A shell consists of all those states with the same value of n, the principal quantum number. A subshell groups all the states within one ...
  12. [12]
    30.3 Bohr's Theory of the Hydrogen Atom – College Physics
    The planetary model of the atom, as modified by Bohr, has the orbits of the electrons quantized. Only certain orbits are allowed, explaining why atomic spectra ...
  13. [13]
    Ionic Compounds | manoa.hawaii.edu/ExploringOurFluidEarth
    The electron shell is the region that the electrons travel in (see Fig. 2.21). Electron shells are labeled with numbers 1 through 7. Each shell holds an ...
  14. [14]
    Electron Configurations & The Periodic Table - MSU chemistry
    The highest occupied electron shell is called the valence shell, and the electrons occupying this shell are called valence electrons. The chemical ...
  15. [15]
    6.9: Lewis Structures of Ionic Compounds – Electrons Transferred
    The tendency to form species that have eight electrons in the valence shell is called the octet rule. The attraction of oppositely charged ions caused by ...
  16. [16]
    2.2 Sodium chloride - Discovering chemistry - The Open University
    Sodium has the electronic configuration, 1s2 2s2 2p6 3s1 (or [Ne] 3s1) so it has one valence electron. It can acquire a filled outer shell by transferring its ...
  17. [17]
    Hybrid Atomic Orbitals
    The relationship between hybridization and the distribution of electrons in the valence shell of an atom is summarized in the table below. Number of Places ...
  18. [18]
    [PDF] Molecular Geometry & Hybridization of Atomic Orbitals
    Valence-Shell Electron-Pair Repulsion Theory. (VSEPR). Theory based on the idea that pairs of valence electrons in bonded atoms repel one another.
  19. [19]
    Evolution of Atomic Theory – Chemistry - JMU Libraries Pressbooks
    In 1904, Thomson proposed the “plum pudding” model of atoms, which described a positively charged mass with an equal amount of negative charge in the form ...
  20. [20]
    [PDF] LXXIX. The scattering of α and β particles by matter and the structure ...
    Geiger*, who found that the distribution for particles deflected between 30 and 150 from a thin gold-foil was in substantial agreement with the theory. A more ...Missing: nuclear | Show results with:nuclear
  21. [21]
    The Bohr Model – Chemistry
    If classical electromagnetic theory is applied, then the Rutherford atom would emit electromagnetic radiation of continually increasing frequency (contrary ...Missing: instability | Show results with:instability
  22. [22]
    The Bohr Atom - Galileo
    Bohr concluded that in an atom in its natural rest state, the electron must be in a special orbit, he called it a "stationary state" to which the usual rules ...Missing: original | Show results with:original
  23. [23]
    Physics 48 (Quantum Mechanics) Home Page, Spring 2009
    Wrapping up the historical prelude: Bohr's atom. Bohr postulates angular momentum quantized in multiples of hbar. Interpreting (discrete) energy level ...
  24. [24]
    Emission Spectrum of Hydrogen
    The Bohr model was based on the following assumptions. The electron in a hydrogen atom travels around the nucleus in a circular orbit.
  25. [25]
  26. [26]
  27. [27]
    A Primer on Quantum Numbers and Spectroscopic Notation
    ... electron with principal quantum number n. The n=1 levels can contain only 2 electrons. This level is called the 1s orbit or the K shell (shells with n=1,2,3 ...
  28. [28]
    The hydrogen atom
    The energy levels scale with Z2, i.e. En = -Z2*13.6 eV/n2. It takes more energy to remove an electron from the nucleus, because the attractive force that must ...
  29. [29]
    [PDF] 1. Physical Constants - Particle Data Group
    Bohr radius (mnucleus = ∞) a∞ = 4π 0h. 2/mee2 = reα−2. 0.529 177 210 903(80) ... 1 Å ≡ 0.1 nm. 1 dyne ≡ 10−5 N. (1 kg)c2 = 5.609 588 603 ... × 1035eV ...
  30. [30]
    Atomic Orbitals and Their Energies
    This peak corresponds to the most probable radius for the electron, 52.9 pm, which is exactly the radius predicted by Bohr's model of the hydrogen atom. For the ...
  31. [31]
    Orbital angular momentum is quantized.
    The possible values that we can measure for the square of the magnitude of the orbital angular momentum are L2 = l(l + 1)ħ2. It is customary to give L2, and not ...
  32. [32]
    [PDF] 221B Lecture Notes
    However, in multi-electron atoms, the degeneracy is lifted by the Coulomb repulsion among electrons, which is a much larger effect. As an example, consider ...
  33. [33]
    [PDF] 99 Chapter 25: Atomic Structure Why can only two electrons occupy ...
    Aug 25, 2021 · The wave function for a multi-electron atom is approximated as the product of one-electron orbitals, which is called the orbital approximation.
  34. [34]
    4.2 Electron shielding and effective nuclear charge - UCF Pressbooks
    Electrons in different orbitals have different electron densities around the nucleus. In other words, penetration depends on the shell (n) and subshell (l).
  35. [35]
    2.6: Slater's Rules - Chemistry LibreTexts
    Sep 5, 2025 · Slater's rules allow you to estimate the effective nuclear charge \(Z_{eff}\) from the real number of protons in the nucleus and the effective ...Slater's Rules · Calculating S · Step B · Step A
  36. [36]
    Quantum Numbers and Atomic Energy Levels - HyperPhysics
    This dependence upon the orbital quantum number of a single excited state electron is accounted for by the penetration of the wavefunction and can be clearly ...
  37. [37]
    Spin Quantum Number - Chemistry LibreTexts
    Jan 29, 2023 · Each orbital can only hold two electrons. One electron will have a +1/2 spin and the other will have a -1/2 spin. Electrons like to fill ...
  38. [38]
    Hund's Rules - Chemistry LibreTexts
    Jan 29, 2023 · Hund's rule states that: Every orbital in a sublevel is singly occupied before any orbital is doubly occupied.
  39. [39]
    6.4 Electronic Structure of Atoms (Electron Configurations) - OpenStax
    Feb 14, 2019 · This subshell is filled to its capacity with 10 electrons (remember that for l = 2 [d orbitals], there are 2l + 1 = 5 values of ml, meaning that ...
  40. [40]
    The Distribution of Electrons among Atomic Levels - chemteam.info
    The Distribution of Electrons among Atomic Levels. Edmund C. Stoner Philosophical Magazine October 1924. Series 6, Volume 48, No. 286 p. 719 - 736.Missing: Aufbau | Show results with:Aufbau
  41. [41]
  42. [42]
    [PDF] Chapter 7 Electronic Configurations and the Properties of Atoms 1
    When electron configurations for multielectron species are written, noble gas notation is often used to represent filled shells (these filled shells are also ...<|control11|><|separator|>
  43. [43]
    6.6: Electron Configurations and the Periodic Table
    The electrons in the highest-numbered shell, plus any electrons in the last unfilled subshell, are called valence electrons; the highest-numbered shell is ...
  44. [44]
    Valence Electrons - The Covalent Bond
    The electrons in the outermost shell are the valence electrons -- the electrons on an atom that can be gained or lost in a chemical reaction.
  45. [45]
    Ionization Energy and Electron Affinity
    In general, the first ionization energy increases as we go from left to right across a row of the periodic table. The first ionization energy decreases as we go ...
  46. [46]
    Property Trends in the Periodic Table - Chemistry 301
    The noble gas configurations are the filled shells and represent the lowest energy most stable configurations. After this, the next electron is forced into a ...
  47. [47]
    Electronic configurations of the elements - math NIST
    Electronic configurations are ground configurations of neutral elements, using notation [X] for noble gas subshells. Data is from NIST, and some configurations ...
  48. [48]
    Atomic Reference Data for Electronic Structure Calculations
    Oct 15, 2015 · The data includes ground electronic configurations of neutral elements, from the NIST Atomic Physics Division, and some first cation  ...
  49. [49]
    [PDF] 5. Electronic Structure of the Elements - Particle Data Group
    May 31, 2024 · The electron configuration for, say, iron indicates an argon electronic core (see argon) plus six 3d electrons and two 4s electrons. Element.
  50. [50]
    6.6 Building Up the Periodic Table
    Because two electrons can be accommodated per orbital, the number of columns in each block is the same as the maximum electron capacity of the subshell: 2 for ...