Fact-checked by Grok 2 weeks ago

Core electron

A core electron is an electron occupying an inner or low-energy in an atom, distinct from electrons in the outermost . These electrons are tightly bound to the due to their proximity and do not directly participate in chemical bonding or reactions. Core electrons play a pivotal role in atomic structure by providing a , where they reduce the felt by valence electrons through electrostatic repulsion, thereby influencing properties such as , , and across the periodic table. This shielding is most pronounced for inner-shell electrons, as they effectively screen the from outer electrons while experiencing minimal shielding themselves. Although inert in typical bonding, core electrons are central to advanced spectroscopic techniques, particularly X-ray-based methods like () and X-ray absorption near-edge structure (XANES), where their excitation or ejection reveals details about an atom's , coordination environment, and local structure in materials. In XPS, for instance, the binding energies of core electrons are measured to identify elemental composition and chemical shifts. In multi-electron atoms, core electrons fill lower shells (n=1, 2, etc.) according to the , forming stable, closed subshells that contribute to the inert nature of and the overall stability of atomic cores. Their relativistic effects become significant in heavy elements, altering orbital energies and influencing phenomena like the .

Basic Concepts

Definition

Core electrons are the electrons in an that occupy the inner atomic shells, specifically the K, L, and M shells, which are tightly bound to the due to their low levels and proximity. These electrons, corresponding to principal quantum numbers n = 1 (K shell), n = 2 (L shell), and n = 3 (M shell), do not participate in chemical bonding because they are shielded from external interactions and remain largely unaffected by neighboring atoms. In , the K shell consists of the 1s orbital, the L shell includes the 2s and 2p orbitals, and the M shell encompasses the 3s, 3p, and 3d orbitals; for lighter elements (up to atomic number around 20), core electrons typically fill orbitals with n < 4. This assignment is based on orbital theory, where electrons occupy discrete energy levels around the nucleus. Core electrons thus form a stable, closed-shell structure that defines the atomic core. A key function of core electrons is to screen the positively charged nucleus from outer electrons, thereby reducing the effective nuclear charge (Z_{\text{eff}}) felt by valence electrons; Slater's rules provide a qualitative approximation for this shielding, treating core electrons in inner shells as contributing nearly fully (0.85–1.00) to the shielding constant for electrons in the same group or outer groups. For instance, in the neon atom (atomic number 10), the core electrons are $1s^2, which shield the valence electrons $2s^2 2p^6, forming a complete noble gas configuration that serves as the core for elements with higher atomic numbers. In contrast, for heavier elements like gold (atomic number 79, configuration [Xe] 4f^{14} 5d^{10} 6s^1), the core includes electrons up to the 4f orbitals, illustrating how the definition expands with increasing atomic number to encompass more inner shells.

Distinction from Valence Electrons

Core electrons are distinguished from valence electrons primarily by their position in the atomic structure and their involvement in chemical processes. Valence electrons occupy the outermost electron shell of an atom, typically those in the highest principal quantum number (n) or unfilled inner shells for transition metals, and are directly responsible for chemical bonding and the determination of an element's chemical properties. In contrast, core electrons reside in the inner shells closer to the nucleus and do not participate in bonding due to their strong attraction to the nuclear charge. A key physical distinction lies in their binding energies, which reflect the energy required to remove these electrons from the atom. Core electrons exhibit significantly higher binding energies than valence electrons, typically ranging from tens of eV to several keV depending on the shell and atomic number, due to their proximity to the nucleus and effective nuclear charge; for instance, the 1s core electron in a has a binding energy of approximately 284 eV. Valence electrons, however, have much lower binding energies, typically less than 20 eV, corresponding to their ionization potentials, which for elements like carbon range from about 11 eV for the first valence electron. This stark energy difference—often orders of magnitude—makes core electrons tightly bound and stable under normal chemical conditions, while valence electrons are more easily excited or removed. Functionally, core electrons remain inert with respect to chemical reactivity, serving instead as a stable foundation that screens the nucleus and influences atomic spectra through high-energy transitions, whereas valence electrons govern an element's position in the , its reactivity, and bonding behavior. For example, in sodium (Na), the electron configuration is 1s² 2s² 2p⁶ 3s¹, where the 1s² 2s² 2p⁶ electrons form the core and are chemically inert, while the single 3s¹ valence electron is readily lost to form the Na⁺ ion, exemplifying its role in ionic bonding. This separation underscores how valence electrons dictate periodicity and chemical trends, with the atomic core—comprising the nucleus and core electrons—providing a passive structural backbone.

Theoretical Framework

Orbital Theory

The quantum mechanical description of core electrons begins with the Schrödinger equation for hydrogen-like atoms, which models a single electron orbiting a nucleus of charge Ze. The time-independent Schrödinger equation in spherical coordinates is given by -\frac{\hbar^2}{2m} \nabla^2 \psi(r, \theta, \phi) - \frac{Z e^2}{4\pi \epsilon_0 r} \psi(r, \theta, \phi) = E \psi(r, \theta, \phi), where \psi(r, \theta, \phi) is the wave function, m is the electron mass, and E is the energy eigenvalue. The solutions, known as atomic orbitals, take the separable form \psi_{n,l,m_l}(r, \theta, \phi) = R_{n,l}(r) Y_{l,m_l}(\theta, \phi), with the radial function R_{n,l}(r) describing the electron's distance from the nucleus and the angular part Y_{l,m_l}(\theta, \phi) given by spherical harmonics. These orbitals form the basis for understanding core electrons, which occupy the lowest principal quantum number n shells closest to the nucleus. Core orbitals are characterized by four quantum numbers: the principal quantum number n (a positive integer determining the shell), the azimuthal quantum number l (ranging from 0 to n-1, with l=0 for s, l=1 for p in inner shells), the magnetic quantum number m_l (from -l to +l), and the spin quantum number m_s (\pm 1/2). The Pauli exclusion principle states that no two electrons in an atom can share the same set of all four quantum numbers, limiting each orbital to a maximum of two electrons with opposite spins and restricting subshell occupancy to $2(2l + 1) electrons (e.g., 2 for 1s, 6 for 2p). In core regions, low n and l values dominate, such as the 1s orbital (n=1, l=0), which is fully occupied in atoms beyond hydrogen. The radial probability distribution $4\pi r^2 |R_{n,l}(r)|^2 for core orbitals shows a high probability density near the nucleus. For the 1s orbital in helium, a representative core electron example, this distribution peaks at approximately 0.13 Å from the nucleus, reflecting the strong nuclear attraction and small spatial extent of inner-shell electrons. This confinement contrasts with valence orbitals, which extend farther out. For multi-electron atoms, the independent-electron approximation of hydrogen-like orbitals is extended using the , which solves self-consistent field equations to account for electron-electron interactions. In this approach, the many-electron wave function is approximated as a of single-particle orbitals, and the effective potential for each electron includes nuclear attraction plus averaged Coulomb and exchange terms from other electrons. Core orbitals, being innermost, experience a nearly bare nuclear potential due to minimal shielding from outer electrons, leading to tight binding. A key feature distinguishing core orbitals is the penetration effect, where s-electrons (l=0) have radial wave functions that extend closer to the nucleus compared to p- or higher-l orbitals in the same shell, as their probability density does not vanish at r=0. This penetration reduces shielding by inner electrons, increasing the effective nuclear charge Z_{\text{eff}} and resulting in higher binding energies for core s-electrons relative to non-penetrating orbitals. For instance, in a given shell, the energy ordering follows E_s < E_p < E_d, with core s-orbitals exhibiting the most negative (tightest bound) energies.

Atomic Core

In multi-electron atoms, the atomic core comprises the nucleus, which contains Z protons, and the core electrons occupying the inner electron shells. These core electrons are tightly bound to the nucleus and do not participate in chemical bonding. The presence of core electrons reduces the net positive charge experienced by outer electrons through screening, resulting in an effective nuclear charge given by Z_{\text{eff}} = Z - \sigma, where \sigma is the screening constant that accounts for the shielding effect of the inner electrons. For valence electrons, the atomic core functions as a pseudo-nucleus, with its effective charge influencing the behavior of the outer electrons. In alkali metals such as lithium and sodium, the core electrons provide nearly complete screening of the nuclear charge, so the single valence electron perceives an effective charge of approximately +1, akin to the core acting as a point-like positive charge similar to a hydrogen nucleus. The atomic core is characterized by a small radius on the order of a few picometers and exceptionally high electron density near the nucleus, reflecting the compact nature of the inner orbitals. In noble gases like helium, where all electrons occupy core-like orbitals, the core encompasses the entire atom, with helium exhibiting an atomic radius of about 31 pm and concentrated electron density in its 1s shell./Descriptive_Chemistry/Elements_Organized_by_Block/2_p-Block_Elements/Group_18%3A_The_Noble_Gases/1Group_18%3A_Properties_of_Nobel_Gases) This high density contributes to the chemical inertness of noble gases by stabilizing the filled inner shells. Core electrons play a key role in periodic trends, particularly the inert pair effect observed in heavier p-block elements such as thallium and lead. Here, the inner d- and f-orbital core electrons provide incomplete shielding, leading to a higher effective nuclear charge on the outermost ns electrons, which become more tightly bound and less available for bonding, favoring lower oxidation states./08%3A_Chemistry_of_the_Main_Group_Elements/8.06%3A_Group_13_(and_a_note_on_the_post-transition_metals)/8.6.02%3A_Heavier_Elements_of_Group_13_and_the_Inert_Pair_Effect)

Physical Phenomena

Relativistic Effects

In high atomic number (high-Z) atoms, core electrons experience significant relativistic effects due to their high velocities near the nucleus, which approach fractions of the speed of light comparable to 0.1c or higher for inner shells. For instance, the 1s electrons in atoms like achieve average velocities of approximately (Z/137)c ≈ 0.58c, where the factor 137 arises from the inverse of the . This relativistic motion increases the effective electron mass according to the Lorentz factor, m = γ m_0 with γ = 1/√(1 - v²/c²), leading to stronger binding and altered orbital characteristics compared to non-relativistic descriptions. These kinematical effects are most pronounced for s electrons, which have maximum probability density at the nucleus, and become increasingly important for Z > 50. The fundamental theoretical framework for incorporating into atomic structure is provided by the , which describes the relativistic of electrons in the field of the . For hydrogen-like atoms, the exact energy levels derived from the are given by E_{n j} = m c^2 \left[ 1 + \left( \frac{Z \alpha}{n - (j + 1/2) + \sqrt{(j + 1/2)^2 - (Z \alpha)^2}} \right)^2 \right]^{-1/2}, where m is the electron rest mass, c is the , Z is the , n is the principal , j is the , and α = e²/(4πε₀ ħ c) ≈ 1/137 is the . This formula accounts for spin-orbit coupling and predicts splittings that match experimental observations in heavy atoms far better than non-relativistic Schrödinger-based models. A key consequence of these relativistic effects is the of s orbitals, such as the 1s, where the expectation value of the ⟨r⟩ is reduced relative to the non-relativistic scaling of ∝ 1/Z², due to the increased effective mass and penetration closer to the . In heavier elements, this leads to inverted energy orderings in shell splittings; for example, the relativistic stabilization of orbitals relative to 5d causes a narrowed 5d-6s gap in (Z=79), shifting absorption to wavelengths and resulting in its characteristic yellow color. Similarly, in mercury, the ² core is strongly stabilized (with the 6s orbital contracted by ~23% relativistically), enhancing the and contributing to mercury's liquid state at by weakening . These effects also amplify the across elements 57La to 71Lu, as progressive 4f filling combined with relativistic contraction reduces atomic radii more sharply than expected non-relativistically.

Electron Transitions

Core-level electron transitions involve the excitation or relaxation of electrons in inner shells, such as the K-shell (1s orbital) to higher shells like the L-shell (2p orbitals), producing or spectra. These transitions occur when an incident or particle ejects a core , creating a "core hole," which is subsequently filled by an electron from a higher shell, releasing energy as an in processes. The frequencies of these emitted X-rays follow , which relates the square root of the frequency ν to the Z: \sqrt{\nu} = (Z - b) \sqrt{ c R \left( \frac{1}{n_1^2} - \frac{1}{n_2^2} \right) }, where c is the , R is the , b is a screening constant (typically ≈1 for K-shell transitions), and n_1 and n_2 are the principal quantum numbers of the initial and final shells (e.g., n_1=1 for K-shell, n_2=2 for L-shell). This empirical relation, derived from experimental spectra of elements, enables precise identification of atomic numbers through line positions. Quantum mechanical selection rules govern the allowed transitions: the change in orbital angular momentum quantum number must satisfy Δl = ±1, and for total angular momentum, Δj = 0, ±1 (with the restriction that j=0 to j=0 is forbidden). Transitions violating these rules, known as forbidden transitions, occur weakly via higher-order multipole interactions and contribute negligibly to observed spectra intensities. In (), core electron transitions are probed by measuring the of photoejected , with binding energies calculated as E_b = hν - E_kinetic, where hν is the incident photon energy and E_kinetic is the measured of the emitted electron (neglecting corrections for simplicity in vacuum measurements). This technique determines core-level binding energies, which are element-specific and sensitive to chemical environment, enabling surface elemental and chemical state analysis. An alternative decay pathway for core holes is the , where a fills the core vacancy, and the released energy ejects another as a secondary , rather than emitting a . This non-radiative process dominates in lighter elements or when emission yields are low, providing complementary information on local electronic structure in . A representative example is the Kα emission line in (Z=29), arising from a 2p to 1s transition, with energies of approximately 8028 eV (Kα2) and 8048 eV (Kα1), often averaged to ~8 keV for practical applications. These lines are widely used in for non-destructive in and .

References

  1. [1]
    1.9B: Valence and Core Electrons - Chemistry LibreTexts
    Jan 15, 2023 · Valence electrons occupy the outermost shell or highest energy level of an atom while core electrons are those occupying the innermost shell or lowest energy ...
  2. [2]
    3.4: Core and Valence Electrons - Chemistry LibreTexts
    Jul 17, 2023 · Valence shell electrons (or, more simply, the valence electrons) are the electrons in the highest-numbered shell, or valence shell, while core ...
  3. [3]
    2.2.4: Shielding - Chemistry LibreTexts
    May 14, 2024 · Core electrons shield valence electrons, but valence electrons have little effect on the Z* of core electrons. The ability to shield, and be ...
  4. [4]
    6.18: Electron Shielding - Chemistry LibreTexts
    Mar 20, 2025 · Electron shielding is the blocking of valence shell electron attraction by the nucleus, due to the presence of inner-shell electrons.
  5. [5]
    41. Core and Valence Electrons, Shielding, Zeff (M7Q8)
    Core electrons are adept at shielding, while electrons in the same valence shell do not block the nuclear attraction experienced by each other as efficiently.
  6. [6]
    XANES - Chemistry LibreTexts
    Jan 29, 2023 · As a core electron absorbs an X-ray photon and is ejected from its core shell, the absence of the electron left in the core shell leads to a ...
  7. [7]
    X-ray Photoelectron Spectroscopy (XPS)
    In X-ray Absorption spectroscopy, a core electron is excited into unocupied atomic/molecular orbitals above the Fermi level. XAS is divided into two regimes; ...
  8. [8]
    a short introduction to atomic structure
    There is also another notation that is commonly used in X-ray spectroscopy. Shells with n = 1, 2, 3, 4, 5, 6 and 7 are indicated with K, L, M, N, O, P, Q. A ...
  9. [9]
    A Primer on Quantum Numbers and Spectroscopic Notation
    The n=1 levels can contain only 2 electrons. This level is called the 1s orbit or the K shell (shells with n=1,2,3,4,5,6,7 are called the K, L, M, N, O, P, Q ...
  10. [10]
    [PDF] Chemical Bonding Michael Morse, University of Utah morse@chem ...
    Core orbitals-. No overlap, too close to the nucleus; no contribution to bonding. Much lower in energy than shown! Valence orbitals-. Responsible for chemical ...
  11. [11]
    CHEM 101 - Atoms: An introduction
    Oct 23, 2023 · The core electrons are those in the innermost shells and are found for the most part close to the nucleus. The electrons in the outermost shell ...
  12. [12]
    Slater's Rules
    For reason #2 the inner core electrons shield the valence electron from the nucleus so the outer most electron only experiences an effective nuclear charge.
  13. [13]
    [PDF] 99 Chapter 25: Atomic Structure Why can only two electrons occupy ...
    Aug 25, 2021 · Slater's Rules are used to give approximations to the effective nuclear charge of the valence electrons. The shielding efficiency of ...<|control11|><|separator|>
  14. [14]
    Electron Configurations & The Periodic Table - MSU chemistry
    In the third period of the table, the atoms all have a neon-like core of 10 electrons, and shell #3 is occupied progressively with eight electrons, starting ...
  15. [15]
    Electron Configurations - EdTech Books
    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).
  16. [16]
  17. [17]
    Lewis1
    Core electrons are electrons which are held close to the nucleus and do not take part in chemical bonding. Valence electrons are electrons that are present in ...
  18. [18]
    Section 1.1 ELECTRON BINDING ENERGIES - X-Ray Data Booklet
    Table 1-1 gives the electron binding energies for the elements in their natural forms. A PDF version of this table is also available.
  19. [19]
    WebElements Periodic Table » Carbon » properties of free atoms
    The electron affinity of carbon is 153.9 1.262118(20) eV kJ mol‑1. The ionisation energies of carbon are given below. Ionisation energies of carbon. Ionisation ...
  20. [20]
    Core Level Spectroscopy
    electron binding energy = photon energy - kinetic energy of the emitted electron - workfunction. Since binding energies of core electrons are characteristic ...
  21. [21]
    19 The Hydrogen Atom and The Periodic Table - Feynman Lectures
    19 The Hydrogen Atom and The Periodic Table. 19–1Schrödinger's equation for the hydrogen atom. The most dramatic success in the history of the quantum ...<|separator|>
  22. [22]
    Atomic Spectros. - Atomic States, Shells, Configs - NIST
    The Pauli exclusion principle prohibits atomic states having two electrons with all four quantum numbers the same. Thus the maximum number of equivalent ...
  23. [23]
    [PDF] B3.1 Quantum structural methods for atoms and molecules
    the nucleus, the maximum of the 1s orbital's radial probability density. The ... distance from the He nucleus is 0.13 Å (the peak of the 1s orbital's ...
  24. [24]
    [PDF] Elements of Atomic Structure in Multi-Electron Atoms
    To summarize, for a given atom, Hartree-Fock theory, combined with averaging of the potentials, gives us an electron configuration; a set of self-consistent ...
  25. [25]
    [PDF] Principles of Chemical Science - MIT OpenCourseWare
    Sep 17, 2014 · Thus in a multi-electron atom, effects of penetration and shielding give rise to the order of energies of orbitals in a given shell of s < p < d ...
  26. [26]
    Effective Nuclear Charge - GeeksforGeeks
    Jul 23, 2025 · Effective Nuclear Charge is the net positive charge experienced by an electron in a multi-electron atom, also known as Core Charge.
  27. [27]
    [PDF] CHEM 115 - UMass Boston
    number of core electrons = 10. 11 protons and some neutrons. (charge: +11). All complete inner shells of electrons. (charge: -10). CORE. Net +1 charge. (Note: ...
  28. [28]
    Atomic Radius of the Elements - Photographic Periodic Table
    Atomic Radius of the elements ; Helium, 31 pm, Cadmium ; Neon, 38 pm, Silver ; Fluorine, 42 pm, Chromium ; Oxygen, 48 pm, Lithium ; Hydrogen, 53 pm, Palladium ...
  29. [29]
    What is the inert pair effect? - Chemistry Stack Exchange
    Feb 25, 2014 · The inert pair effect describes the preference of late p-block elements (elements of the 3rd to 6th main group, starting from the 4th period)
  30. [30]
    Atomic Spectros. - Spectral Lines - NIST
    Selection rules for discrete transitions. Electric dipole (E1) ("allowed"), Magnetic dipole (M1) ("forbidden"), Electric quadrupole (E2) ("forbidden") ...
  31. [31]
    Henry Moseley, X-ray spectroscopy and the periodic table - Journals
    Aug 17, 2020 · This paper provides an introduction to Moseley and his experiments and then traces attempts to 'discover' missing elements by X-ray spectroscopy.Moseley's papers on 'High... · Mopping up X-ray spectra in... · Beyond Z = 100
  32. [32]
    [PDF] High-Frequency Spectra of the Elements. 703 - MIT
    The quantitative measurements of. Moseley and Darwin on the reflexion of the general radiation must have been little affected by these fringes, as the incident ...
  33. [33]
    [PDF] The NIST X-ray photoelectron spectroscopy (XPS) database
    The NIST XPS database contains data on binding and kinetic energy of sample electrons from all elements, compiled from 13,200 records of 800 papers.
  34. [34]
    Auger processes in the 21st century - PMC - PubMed Central
    A valence electron can fill the core hole liberating enough energy for the emission of a second electron, an Auger electron. The kinetic energy of the Auger ...Missing: seminal | Show results with:seminal
  35. [35]
    [PDF] Table 1-2. Energies of x-ray emission lines
    Photon energies, in electron volts, of principal K-, L-, and M-shell emission lines. ... 29 Cu. 8,047.78 8,027.83. 8,905.29. 929.7. 929.7. 949.8. 30 Zn. 8,638.86 ...
  36. [36]
    High-precision measurement of the X-ray Cu K-alpha spectrum | NIST
    May 12, 2017 · In this measurement, the region from 8000 to 8100 eV has been covered with a highly precise angular scale, and well-defined system efficiency.