Fact-checked by Grok 2 weeks ago

Electron affinity

Electron affinity is the amount of energy released when an electron is added to a neutral atom or molecule in the gas phase to form a singly charged negative ion, typically measured in kilojoules per mole (kJ/mol) or electron volts (eV). This property quantifies an atom's tendency to gain an electron, with more negative values indicating greater stability for the resulting anion due to exothermic energy release; for instance, chlorine has an electron affinity of -349 kJ/mol, while noble gases like helium have near-zero or positive values, reflecting their reluctance to accept electrons. In the periodic table, electron affinity generally becomes more negative (increases in magnitude) from left to right across a , driven by increasing that pulls electrons closer to the despite similar atomic sizes. It decreases down a group, as larger atomic radii and increased electron shielding reduce the attraction for additional electrons; exhibit the highest affinities (e.g., at -328 kJ/mol), while metals have lower values (e.g., at -60 kJ/mol). Exceptions occur, such as nitrogen's lower affinity than carbon due to electron repulsion in a half-filled p subshell, highlighting the role of stability. Electron affinity is crucial for understanding chemical reactivity, particularly in ion formation and bonding, as it influences the ease with which nonmetals form anions in ionic compounds and contributes to electronegativity trends. It also plays a key role in predicting energies in solids and the behavior of elements in reactions, with high-affinity elements like being strong oxidizing agents. In advanced applications, such as physics, electron affinity determines band structures and mobility in materials like .

Fundamentals

Definition

Electron affinity refers to the energy released during the thermodynamic process of adding an electron to a neutral atom or molecule in the gas phase, represented as X(g) + e⁻ → X⁻(g), where the process is typically exothermic for most elements. This quantity, denoted as E_{ea}, quantifies the stability gained by forming the anion and is defined as the difference between the energy of the initial state (neutral species plus free electron) and the final state (anion), such that E_{ea} = E(\text{initial}) - E(\text{final}). Note that conventions for the sign of electron affinity vary; some sources report it as the negative energy change (ΔE < 0 for exothermic), while others (including this section and NIST) report the positive energy released. Equivalently, E_{ea} = -\Delta E(\text{attachment}), where \Delta E(\text{attachment}) is the energy change for the attachment process; this convention emphasizes E_{ea} as positive when energy is released, distinguishing it from the attachment energy, which is often reported as the negative value of this change. Electron affinity values are conventionally expressed in kilojoules per mole (kJ/mol) or electronvolts (eV), with the conversion factor being $1 \, \text{eV} \approx 96.485 \, \text{kJ/mol}. The reverse process requires energy equal to the electron affinity value to detach the from the anion: X⁻(g) → X(g) + e⁻. This is distinct from ionization energy, which measures the energy required to remove an electron from a neutral atom to form a cation.

Relation to Other Atomic Properties

Electron affinity (EA) is closely related to ionization energy (IE), representing the complementary process in electron transfer. Whereas IE quantifies the energy required to remove an electron from a neutral gaseous atom to form a cation, EA measures the energy released when a neutral gaseous atom gains an electron to form an anion. These properties together describe an atom's capacity for electron loss or gain, and their combination appears in models for molecular dissociation; for instance, in certain diatomic ions, the sum of IE and EA approximates the energy associated with dissociation into neutral components, incorporating Coulombic interactions. A key interconnection arises in the Mulliken electronegativity scale, where electronegativity \chi is defined as the average of the ionization energy and : \chi = \frac{\mathrm{IE} + \mathrm{EA}}{2} This formulation, proposed by in 1934, provided the first quantitative link between these atomic properties and an atom's ability to attract electrons in chemical bonds, establishing electronegativity on an absolute energy scale. Electronegativity scales further illustrate EA's role. The Pauling scale, based on bond dissociation energies, indirectly incorporates EA through its influence on bond polarity and strength, as higher EA contributes to greater electron attraction in heteronuclear bonds. In contrast, the Allred-Rochow scale explicitly ties electronegativity to the effective nuclear charge Z_\mathrm{eff}, calculated as \chi = 0.359 \frac{Z_\mathrm{eff}}{r^2} + 0.744 (in Pauling units, with r as covalent radius), which directly affects EA by determining the attraction for incoming electrons. Electron gain enthalpy, often denoted \Delta_\mathrm{eg} [H](/page/H+), is closely related but distinct from EA in thermodynamic contexts. While EA typically refers to the energy change at 0 K for the process \ce{X(g) + e^- -> X^-(g)}, electron gain enthalpy accounts for the enthalpy change \Delta [H](/page/H+) under typical conditions (e.g., 298 K), where \Delta_\mathrm{eg} [H](/page/H+) \approx -\mathrm{EA} since are small and often negligible (~2-3 /). The terms are sometimes used synonymously, particularly when are negligible, but \Delta_\mathrm{eg} [H](/page/H+) emphasizes the enthalpic in thermochemical cycles like the Born-Haber .

Measurement and Conventions

Experimental Methods

The primary experimental methods for measuring electron affinities in gas-phase atomic and molecular systems rely on spectroscopic and collisional techniques that probe the energy required to detach an from a negative or the for anion formation. These approaches have evolved significantly since the mid-20th century, with modern variants achieving resolutions down to a few microelectronvolts, enabling precise determination of electron binding energies. Key techniques include , photodetachment, electron impact excitation, and equilibrium studies using ion traps or flow reactors. Photoelectron spectroscopy, particularly negative ion photoelectron spectroscopy (PES), is one of the most widely used methods for directly measuring electron affinities. In this technique, mass-selected negative ions are irradiated with a fixed-frequency laser, and the kinetic energy distribution of the detached photoelectrons is analyzed using a hemispherical electron energy analyzer or time-of-flight setup. The electron affinity EA is derived from the relation EA = h\nu - KE, where h\nu is the photon energy and KE is the maximum kinetic energy of the ejected electrons corresponding to the adiabatic transition from the anion to the neutral ground state. Advanced variants, such as slow electron velocity-map imaging (SEVI), enhance resolution by focusing low-energy electrons, achieving uncertainties as low as 1-5 meV for species like halogens and transition metals. For example, SEVI has been applied to determine the electron affinity of lead with a precision of 16 μeV. This method excels for systems with electron affinities up to about 3.5 eV but requires careful correction for vibrational hot bands and rotational structure. Laser photodetachment threshold spectroscopy provides high-precision measurements by scanning a tunable laser to identify the photon energy threshold for electron detachment from the anion, where EA = h\nu_{\text{threshold}} at zero kinetic energy. Typically performed with mid-infrared or visible tunable lasers in an ion beam or trap setup, this method minimizes systematic errors through calibration with known standards like the hydrogen anion. It has yielded values such as 0.754195 ± 0.000019 eV for hydrogen and 3.612720 ± 0.000044 eV for chlorine. Recent innovations, including multi-reflection time-of-flight (MR-ToF) devices, confine ions for multiple laser interactions, boosting sensitivity by three orders of magnitude and enabling measurements with minimal sample quantities (e.g., ~6100 ions for chlorine), which is particularly advantageous for rare elements like actinides. Challenges include identifying the exact threshold amid weak signals or overlapping electronic states. Electron impact methods involve colliding low-energy with neutral atoms or molecules to form anions and measuring the for attachment, which approximates the negative of the electron affinity. These experiments often use crossed-beam setups or mass spectrometers to detect anion yields as a of incident electron energy, with thresholds extracted from onset curves. While historically important for early measurements (e.g., at 1.07 ± 0.07 eV), this approach suffers from lower precision, typically ±0.1-0.2 eV, due to background and difficulties in resolving near-threshold processes. It is less favored today for high-accuracy work but remains useful for surveying molecular systems with moderate electron affinities. Equilibrium studies derive electron affinities indirectly through thermochemical cycles involving gas-phase -molecule , often measured in ion cyclotron resonance (ICR) mass spectrometers or flowing apparatuses. In ICR, constants for like A^- + B \rightleftharpoons A + B^- are determined from forward and reverse ratios at controlled temperatures, yielding \Delta H values related to EA(A) - EA(B). Flowing techniques similarly monitor intensities in a carrier gas to establish , as demonstrated for (1.1 ± 0.1 eV). These methods provide reliable data for species where direct photodetachment is challenging, with accuracies around ±0.05-0.1 eV when anchored to known standards, though they require precise and auxiliary bond dissociation energies. Anion zero kinetic energy (ZEKE) spectroscopy represents a high-resolution extension of photodetachment methods, selectively detecting near-zero electrons after a delay to discriminate against prompt photoelectrons. Using pulsed field ionization or velocity-map imaging, it achieves resolutions below 1 meV for vibrational and electronic structure, as applied to anions and clusters like Ge₂⁻. Overall accuracies for electron affinities range from 0.005-0.020 eV in standard PES and photodetachment, improving to sub-meV with ZEKE or MR-ToF, but challenges persist for and elements with near-zero affinities (e.g., uncertainties ~0.1 eV for due to unstable anions and weak signals). Recent advancements as of 2025, such as enhanced ion confinement, address sample scarcity and low detachment cross-sections, expanding applicability to superheavy elements. Results are reported using the standard convention where positive values indicate exothermic electron attachment.

Sign Convention and Units

In the early 20th century, the for electron affinity was inconsistent, with some sources reporting positive values for endothermic processes (where energy is absorbed upon electron attachment), leading to confusion in interpreting data for elements like alkaline earth metals. This ambiguity arose from varying definitions in early spectroscopic and thermodynamic studies, where electron affinity was sometimes equated directly to the energy change without standardized sign usage. The modern convention, as defined by the IUPAC Gold Book, establishes affinity E_\text{ea} as the positive energy required to detach an from the singly charged negative in the gas , corresponding to the process \ce{X^- -> X + e^-}. This is equivalent to E_\text{ea} = -\Delta E for the reverse attachment process \ce{X + e^- -> X^-}, where \Delta E < 0 indicates energy release, ensuring E_\text{ea} > 0 for exothermic attachments that form stable anions. Older texts occasionally used \Delta E directly for attachment, resulting in negative values for exothermic cases, but the IUPAC approach aligns with conventions for consistency. Electron affinity is reported in electronvolts per atom (eV/atom) in atomic physics or kilojoules per mole (kJ/mol) in chemistry, with the conversion factor being $1 \, \text{eV/atom} = 96.485 \, \text{kJ/mol}. Values are often rounded for practical use; for example, the electron affinity of chlorine is 349 kJ/mol, equivalent to approximately 3.61 eV. Thermodynamically, at 0 K approximation, the standard enthalpy of formation for an anion relates to that of the neutral atom by \Delta H_f(\ce{X^-}) = \Delta H_f(\ce{X}) - E_\text{ea}, where E_\text{ea} accounts for the energy released upon electron attachment. This relation facilitates connections between gas-phase ion data and formation enthalpies in thermochemical cycles.

Atomic Electron Affinities

Electron affinity generally becomes more negative across a period from left to right in the periodic table, as the effective nuclear charge experienced by valence electrons rises while the atomic radius decreases, enhancing the attraction for an additional electron. This trend is most pronounced among the p-block elements, where nonmetals exhibit more negative values than s-block metals; for example, in period 2, the value becomes more negative from carbon (-122 kJ/mol) to fluorine (-328 kJ/mol), though group 15 elements like nitrogen show less negative affinities due to stable half-filled subshells. Down a group, electron affinity typically becomes less negative because the increasing and electron shielding by inner shells weaken the nuclear pull on the incoming electron. In the group (17), values become less negative from (-349 kJ/mol) to (-325 kJ/mol) to iodine (-295 kJ/mol), with (-328 kJ/mol) slightly less negative than chlorine owing to interelectronic repulsions in its compact 2p orbitals. Similarly, for the chalcogens (group 16), the affinity becomes more negative from oxygen (-141 kJ/mol) to (-200 kJ/mol) but then less negative to (-195 kJ/mol) and (-190 kJ/mol) due to size effects. exhibit near-zero electron affinities, consistent with their filled shells that resist further electron addition. Key factors influencing these trends include , which inversely correlates with electron affinity as smaller atoms allow stronger electrostatic attraction, and electron configuration stability, where half-filled or fully filled subshells (e.g., in or ) reduce affinity by increasing the energy required to accommodate an extra electron. These patterns are well-reproduced by computational methods such as (DFT), which aligns predictions with experimental data for main group elements, often achieving mean absolute errors below 10 kJ/mol when using . The following table presents selected experimental electron affinity values (in kJ/mol; negative values indicate energy released upon electron attachment) for representative elements in groups 16, 17, and 18:
GroupElementElectron Affinity (kJ/mol)
16 (Chalcogens)O-141
S-200
Se-195
Te-190
17 (Halogens)F-328
Cl-349
Br-325
I-295
18 (Noble Gases)Ne≈0
Ar≈0
Kr≈0

Exceptions and Anomalies

One notable deviation from the expected periodic trend occurs in the , where the electron affinity of (-328 kJ/mol) is less negative than that of (-349 kJ/mol), despite fluorine's higher . This anomaly stems from fluorine's exceptionally small , which causes the added in the F⁻ anion to experience intense repulsion from the tightly packed electrons in the 2p orbitals, destabilizing the ion relative to the larger Cl⁻. Noble gases display particularly striking anomalies, with electron affinities that are generally near zero and often slightly positive (endothermic) due to their filled shells, making anion formation unfavorable in the . For , the value is approximately 0 , though theoretical and experimental studies debate whether it is precisely zero or slightly positive (indicating an unbound ), with no stable anion observed. Heavier like show small binding in metastable configurations, with recent studies (as of 2024) indicating weak attachments in excited electronic states facilitated by advanced laser spectroscopy techniques; remain unbound. In transition metals, electron affinities deviate irregularly from trends due to the variable of d-orbital configurations, often resulting in relatively low (less negative) values. For instance, exhibits an electron affinity of -0.67 eV, attributed to the high of its half-filled 3d⁵ 4s¹ , which requires electronic rearrangement upon adding an to form Cr⁻. Alkali metals represent another class of anomalies, with electron affinities that are notably small (less negative) compared to other groups, reflecting their low and large size. , for example, has an electron affinity of -53 kJ/mol, indicating a small for Na + e⁻ → Na⁻, but the anion is weakly bound due to the weak for an additional in the diffuse 3s orbital. This contrasts sharply with the highly negative affinities in adjacent groups, underscoring the role of ns¹ configurations in limiting anion . Recent advancements using techniques like photoelectron spectroscopy have refined values for noble gases, reducing error margins and emphasizing the role of relativistic effects in heavier atoms.

Molecular Electron Affinities

Differences from Atomic Cases

In molecular systems, electron affinity is profoundly influenced by structural factors such as bond polarity and electron delocalization, which can either enhance or diminish the stability of the resulting anion compared to isolated atoms. For instance, in perfluorocarbons like CF₃, the strong electronegativity of fluorine atoms creates highly polar C-F bonds that withdraw electron density, stabilizing the extra electron in the anion and yielding an electron affinity of 1.82 eV, significantly higher than that of atomic carbon (1.27 eV) due to inductive effects. In contrast, delocalization in diatomic molecules like O₂ leads to a lower electron affinity of 0.451 eV compared to atomic oxygen (1.461 eV), as the added electron forms a weakly bound superoxide ion (O₂⁻) where π* orbital occupancy reduces binding strength despite the molecular framework. A key distinction in molecular electron affinities arises from the vertical and adiabatic processes. The vertical electron affinity measures the energy change for electron attachment without altering the , often approximating the vertical energy from the anion; in contrast, the adiabatic electron affinity accounts for structural relaxation in both the neutral and anion to their geometries, which can differ substantially in polyatomic molecules due to vibrational and conformational adjustments. In many polyatomic species, this relaxation energy leads to adiabatic values that are 0.2–1 eV higher than vertical ones, as the anion's diffuse electron cloud prompts significant geometry changes, such as bond lengthening or angle variations, that stabilize the anion. Molecular anions often exhibit enhanced stability relative to atomic cases through environmental effects like electron scavenging, where the molecular cage or framework traps the added electron. Sulfur hexafluoride (SF₆), for example, has an adiabatic electron affinity of 1.05 eV, enabling it to form a stable SF₆⁻ ion via attachment to antibonding orbitals, a process facilitated by the symmetric octahedral structure that delocalizes the electron and prevents dissociation—contrasting with atomic sulfur's higher isolated affinity of 2.08 eV but lacking such trapping. Perfluorocycloalkanes further illustrate this, with cyclic structures promoting higher affinities (e.g., 0.63 eV for c-C₄F₈) than their acyclic analogs due to constrained geometry that better accommodates the anion's charge distribution. Computing molecular electron affinities poses unique challenges compared to atomic ones, primarily due to the need for basis sets that adequately describe the diffuse, loosely bound orbitals of anions. Standard basis sets often fail to capture the extended nature of the extra , leading to underestimated affinities or artificial instabilities; augmented diffuse functions (e.g., aug-cc-pVTZ) are essential, increasing computational cost by factors of 2–5 for polyatomics while improving accuracy to within 0.1–0.2 of experiment. methods like B3LYP perform well for closed-shell transitions but struggle with open-shell cases or self-interaction errors in delocalized systems, necessitating higher-level approaches such as coupled-cluster theory for reliable predictions in complex molecules like perfluorocarbons.

Selected Examples and Applications

Selected examples of molecular electron affinities illustrate the diversity in stability of molecular anions compared to atomic cases, where values can range from highly positive, indicating stable anions, to negative, signifying unstable ones that autodetach the . For instance, exhibits a negative electron affinity of approximately -1.15 eV, resulting in an unstable that rapidly loses the added due to the molecule's closed-shell aromatic structure. In contrast, (NO), an odd-electron molecule, has a low but positive adiabatic electron affinity of 0.026 eV, allowing formation of a weakly bound anion relevant to reactive intermediates. Halogen molecules like (Cl₂) show a more substantial positive value of 2.38 eV, facilitating stable anion formation due to the electronegative . These electron affinities play critical roles in atmospheric chemistry, particularly electron capture processes. Oxygen's electron affinity of 0.451 eV enables the formation of O₂⁻ ions in the ionosphere, where free electrons attach to O₂ molecules, influencing ionospheric conductivity and plasma dynamics. In organic synthesis, molecular electron affinities determine the feasibility of generating radical anions as reactive intermediates; for example, molecules with positive electron affinities, such as certain nitroaromatics, form persistent radical anions that participate in reduction reactions and coupling processes. In contexts, electron affinity influences the of dyes and polymers, where high electron affinity promotes n-type doping and enhances charge . Studies in the 2020s have explored electron-deficient polymers with electron affinities around 2-3 to improve conductivity in . Recent research on molecular wires highlights how tuning electron affinity to values exceeding 2 enables efficient single-molecule conductance, as demonstrated in alkynylplatinum-based wires for nanoelectronic applications. The following table summarizes electron affinity values for selected common molecules, drawn from the NIST Chemistry WebBook (updated through 2025), emphasizing representative cases across positive and negative regimes.
MoleculeElectron Affinity (eV)Notes
O₂0.451 ± 0.007Adiabatic value for stable atmospheric anion.
NO0.026 ± 0.005Low value for odd-electron species.
Cl₂2.38 ± 0.10Positive, stable anion.
F₂3.01 ± 0.07High value due to electronegativity.
1.33Relevant for carbonyl reactivity.
CO₂-0.60Negative, metastable .
-1.15Negative for aromatic stability.
H₂O≈ -0.8Negative, diffuse anion .

Electron Affinity in Solid State Physics

Definition in Solids

In solid-state physics, the electron affinity \chi of a crystalline solid, particularly a , is defined as the energy difference between the level E_{\text{vac}} and the conduction band minimum E_C, expressed as \chi = E_{\text{vac}} - E_C. This quantity represents the energy released when an is transferred from the level into the lowest unoccupied state at the bottom of the conduction band. Unlike the gas-phase electron affinity for isolated atoms or molecules, which measures the energy change upon adding an electron to a neutral , the solid-state version is inherently dependent on the periodic band structure of the crystal lattice. For instance, the electron affinity of is 4.05 , significantly higher than the 1.39 for an isolated , reflecting the influence of states in the solid. Electron affinities in solids are commonly calculated using (DFT) with slab models, where a periodic of the material is separated by a region to establish the reference vacuum level through the average electrostatic potential; the conduction band edge is then aligned relative to this level. For heterojunctions between different semiconductors, Anderson's rule provides an approximation for the conduction band offset as the difference in the electron affinities of the two materials, assuming alignment of their vacuum levels. In semiconductor physics, electron affinity is conventionally reported as a positive value in electron volts (eV), corresponding to the energy scale of band alignments and interface properties, distinct from some gas-phase chemistry conventions where sign may vary based on endothermic or exothermic processes.

Role in Materials and Devices

In semiconductor devices, electron affinity plays a crucial role in determining the nature of metal-semiconductor contacts, particularly through its influence on Schottky barrier heights. For n-type semiconductors, the Schottky barrier height is given by the difference between the metal work function and the semiconductor's electron affinity, which governs electron injection and rectification behavior. In wide-bandgap materials like gallium nitride (GaN), a relatively high electron affinity of approximately 4.1 eV facilitates efficient electron transport across junctions, enabling high-performance blue light-emitting diodes (LEDs) where low-barrier ohmic contacts are essential for minimizing voltage drops. This property also aids n-type doping strategies, as the alignment of donor levels relative to the conduction band edge—positioned by electron affinity—ensures effective carrier concentration without excessive ionization energies. The electron affinity rule extends to surface physics, where it underpins the formation of heterojunctions and interfaces in devices. In metal-semiconductor contacts, the rule predicts barrier formation based on the semiconductor's electron affinity and the metal's , with the semiconductor work function approximated as the sum of electron affinity, , and position. This relation is vital for optimizing ohmic or rectifying contacts in transistors and sensors, as deviations from ideal alignment due to can lead to Fermi level pinning, but careful material selection leveraging electron affinity minimizes such effects. In emerging materials, electron affinity is key to enhancing charge separation and injection in photovoltaics involving perovskites and two-dimensional (2D) structures. For perovskite solar cells, tuning the electron affinity of transport layers to around 3.8–4.5 eV optimizes band alignment, promoting efficient electron extraction while suppressing recombination, as seen in inverted architectures where high-affinity hole transporters enable fast carrier dynamics. Similarly, graphene's electron affinity near 4.5 eV positions it as an effective electron acceptor in 2D heterostructures, facilitating Schottky-type photovoltaic effects in graphene-semiconductor junctions by lowering barriers for electron injection and improving open-circuit voltages. In 2D-perovskite hybrids, this property enhances interfacial charge transfer, boosting power conversion efficiencies beyond 25% in tandem configurations. Recent advancements in for organic light-emitting diodes (OLEDs) highlight electron affinity's role in charge and device efficiency. Studies propose that organic materials with electron affinities exceeding 3.5 eV may enable trap-free bipolar by reducing electron trapping, potentially allowing single-layer OLEDs with balanced charge injection. Interface engineering, informed by energy level alignment including electron affinity, lowers turn-on voltages to below 3 V in OLEDs, as reported in 2024 reviews of charge injection techniques. Beyond optoelectronics, electron affinity influences applications in and by dictating carrier selectivity and reaction energetics. In polymer-fullerene , higher electron affinities in acceptors like PCBM drive efficient dissociation, contributing to stability against photooxidation and long-term device lifetimes. For , tuning electron affinity in metal oxides, such as Mo-doped cobalt oxides, promotes oxygen evolution reaction (OER) intermediates like *OH adsorption, improving acidic water efficiencies with overpotentials of 348 mV at 10 mA/cm². In birnessite-based electrocatalysts, higher electron affinity correlates with stronger oxidative performance, enabling selective anion adsorption and advancing sustainable fuel production pathways as per 2025 investigations.

References

  1. [1]
    Electron Affinity | Periodic Table of Elements - PubChem
    Electron affinity refers to the energy released when an additional electron is attached to a neutral atom to form a singly charged negative ion.
  2. [2]
    Electron Affinity - Chemistry LibreTexts
    Jan 29, 2023 · Electron affinity is the energy change when a neutral atom gains an electron to form a negative ion, measuring its likelihood of gaining an  ...Missing: authoritative sources
  3. [3]
    6.19: Periodic Trends - Electron Affinity - Chemistry LibreTexts
    Mar 20, 2025 · In general, electron affinities increase (become more negative) from left to right across a period and decrease (become less negative) from top to bottom down ...Missing: authoritative | Show results with:authoritative
  4. [4]
    Ionization Energy and Electron Affinity
    The electron affinity of an element is the energy given off when a neutral atom in the gas phase gains an extra electron to form a negatively charged ion. A ...
  5. [5]
    4.4 Ionization energy and Electron Affinity – Chemistry Fundamentals
    Electron affinity (the energy associated with forming an anion) is more favorable (exothermic) when electrons are placed into lower energy orbitals, closer to ...
  6. [6]
  7. [7]
    Electron Affinity | Research Starters - EBSCO
    Electron affinity is a measure of how strongly a neutral atom or molecule attracts an additional electron, reflecting its ability to form negative ions.
  8. [8]
    Gas-Phase Ion Thermochemistry - the NIST WebBook
    The electron affinity is defined as the negative of the 0 K enthalpy change for the electron attachment reaction: M + e− → M− Δ Hrxn = −EAa. -EA = ΔfH;(M ...
  9. [9]
    [PDF] MASTER^ - OSTI
    The electron affinity EA is defined as the difference in energies of the neutral atom Eo and negative ion states E_ given by. EA=E 0 -E_. (1). EA must be ...<|control11|><|separator|>
  10. [10]
    A New Electroaffinity Scale; Together with Data on Valence States ...
    A new approximate ``absolute'' scale of electronegativity, or electroaffinity, is set up. The absolute electroaffinity on this scale is equal to the average of ...
  11. [11]
    Revisiting Electronegativity and Electronegativity Scales
    Dec 31, 2024 · This scale links the electronegativity coefficient (χ) to the average of the ionization energy I and the electron affinity Ea: χ = 1/2{I + Ea}.Missing: formula | Show results with:formula
  12. [12]
    electron affinity (E01977) - IUPAC
    Energy required to detach an electron from the singly charged negative ion (energy for the process molecule).
  13. [13]
  14. [14]
  15. [15]
    A new action photoelectron spectroscopy for anions - AIP Publishing
    Apr 8, 2005 · The importance of anion ZEKE lies in its ability to provide high resolution vibrational and electronic spectroscopy on anionic mass selected ...
  16. [16]
    Electron Affinities of the Alkaline Earth Metals and the Sign ...
    An argument for a return to the original sign convention for the electron affinity is presented in this paper. ACS Publications. Subjects.
  17. [17]
    Chlorine atom - the NIST WebBook
    Electron affinity determinations ; 3.612709 ± 0.000087, LPD, Trainham, Fletcher, et al., 1987 ; 3.613577 ± 0.000044, LPD, Berzinsh, Gustafsson, et al., 1995 ...
  18. [18]
    Electron Affinity for all the elements in the Periodic Table
    Electron Affinity of the elements ; Neodymium, 50 kJ/mol, Tellurium, 190.2 kJ/mol, Tennessine ; Promethium, 50 kJ/mol, Selenium, 195 kJ/mol, Oganesson ; Samarium ...Missing: NIST noble
  19. [19]
    [PDF] Predicting Electron Affinities with Density Functional Theory
    Sep 6, 2024 · These statistics are determined for the sets of atomic, diatomic, and triatomic electron affinities as well as for the larger set cre- ated by ...Missing: main | Show results with:main
  20. [20]
    Electron affinity of helium | Phys. Rev. A
    Feb 1, 1979 · It is argued that the true electron affinity should not differ from A n r by more than ±0.2 meV. Good agreement with the less-precise ...
  21. [21]
    Noble Gas Anions: An Overview of Strategies and Bonding Motifs
    May 12, 2024 · This review article aims to provide an overview of the strategies employed to prepare noble gas anions under different environments and experimental conditions.
  22. [22]
    Atomic and Molecular Electron Affinities: Photoelectron Experiments ...
    The electron affinity is simply the difference (Figure 1) between the incident photon, hν0, and the KE of the electrons arising from the (0,0) band: EA(R) = hν0 ...
  23. [23]
    The electron affinities of the perfluorocarbons C2Fn, n=1–6
    Nov 22, 1997 · Three measures of neutral-anion separation are reported: the adiabatic electron affinity, the vertical electron affinity, and the vertical ...
  24. [24]
    Oxygen - the NIST WebBook
    Electron affinity determinations ; 0.50 ± 0.20, NBIE, Lacmann and Herschbach, 1970 ; 0.430 ± 0.020, Kine, Pack and Phelps, 1966 ; >0.479998, Endo, Berkowitz, ...Gas phase ion energetics data · References
  25. [25]
    High-precision electron affinity of oxygen - PMC - NIH
    Oct 7, 2022 · The weighted arithmetic mean of the thresholds gives an electron affinity of 1.461,112,972(45) eV for oxygen. The electron affinity is ...
  26. [26]
    Theoretical calculation of the adiabatic electron affinity and vertical ...
    Moreover, the vertical electron affinity (VEA) of all investigated compounds were calculated using the B3LYP/6–31 + G(d,p) level. The substitution effects on ...
  27. [27]
    Sulfur hexafluoride - the NIST WebBook
    Electron affinity determinations ; 1.39 ± 0.13, Kine, Heneghan and Benson, 1983 ; 0.799982, Kine, Lifshitz, 1983 ; 1.20 ± 0.30, N/A, Chase Jr., Curnutt, et al., ...Gas phase ion energetics data · References
  28. [28]
    On the electron affinity of perfluorocycloalkanes and perfluoroalkanes
    There is an increasing body of evidence showing that perfluorocycloalkanes have a higher electron affinity than their open chain analogs, the perfluoroalkanes.Missing: perfluorocarbons | Show results with:perfluorocarbons
  29. [29]
    Efficient Diffuse Basis Sets for Density Functional Theory
    Feb 16, 2010 · Perhaps the most difficult test of the adequacy of the diffuse part of the basis set is the prediction of electron affinity values, because it ...
  30. [30]
    [PDF] Accurate Electron Affinities and Orbital Energies of Anions ... - OSTI
    We assess the accuracy of orbital energies and electron affinities for all three rows of elements in the periodic table (H-Ar) using a variety of theoretical ...
  31. [31]
    From Halogen to Superhalogen Behavior of Organic Molecules ...
    Oct 15, 2015 · Benzene, the classic organic molecule obeying Hückel′s rule of aromaticity, has negative electron affinity (EA), namely −1.15 eV.
  32. [32]
    Nitric oxide - the NIST WebBook
    Proton affinity at 298K ; 526.1 ± 1.3, Kuo, Zhang, et al., 1997, T = 0K; Photoionization of HNO yields DHf(HNO+) from which PA(NO) is calculated at 0 K and 298K.<|separator|>
  33. [33]
    Chlorine - the NIST WebBook
    Electron affinity determinations ; 2.40 ± 0.20, NBIE, Dispert and Lacmann, 1977, B ; 2.32 ± 0.10, Endo, Hughes, Lifschitz, et al., 1973, B.
  34. [34]
    [PDF] Influence of dopant size and electron affinity on the electrical ...
    The study investigates how dopant size and electron affinity affect doping efficiency, electrical conductivity, and thermoelectric properties of conjugated  ...Missing: dyes 2020s
  35. [35]
    Single Molecule Junction of Polynuclear Alkynylplatinum-Based ...
    Feb 8, 2025 · Recent studies have shown that electron-rich di(alkynyl)ruthenium-based wires with diverse bulky ligands can exhibit high conductance ...
  36. [36]
    fluorine - the NIST WebBook
    Electron affinity determinations ; 3.005 ± 0.071, R-A, Wenthold and Squires, 1995, EA fixed at 0K value, not 298K of heat of formation; B ; 3.120 ± 0.070, CIDC ...
  37. [37]
    Carbon monoxide - the NIST WebBook
    Electron affinity determinations. EA (eV), Method, Reference, Comment. 1.32608, R-A, Refaey and Franklin, 1976, G3MP2B3 calculations indicate an EA of ca.-1.6 ...
  38. [38]
    Carbon dioxide - the NIST WebBook
    Electron affinity determinations ; -0.599986, EIAE, Knapp, Echt, et al., 1986, Unbound but in -0.3 eV well, from (CO2)n; B ; -1.60 ± 0.10, NBIE, Compton, ...Gas phase ion energetics data · References
  39. [39]
    Calculated Electron Affininty for H 2 O (Water)
    Calculated Electron Affininty for H2O (Water) ; -3.766 ...
  40. [40]
    Electron Affinity - an overview | ScienceDirect Topics
    ... electron is called the electron affinity. When the negative ion thus formed is stable, energy is released and the electron affinity has a negative sign.
  41. [41]
    Introduction To Semiconductors (all content) - DoITPoMS
    The energy difference labelled as χ between the bottom of the conduction band and infinity is called the electron affinity of the semiconductor. Re-use of ...
  42. [42]
    Basic Parameters of Silicon (Si)
    Silicon has a diamond crystal structure, 5x10^22 atoms/cm^3, density of 2.329 g/cm^3, dielectric constant of 11.7, and electron affinity of 4.05 eV.
  43. [43]
    silicon - the NIST WebBook
    Silicon (Si) has a molecular weight of 28.0855. Its electron affinity is 1.389517 eV and ionization energy is 8.15169 eV.
  44. [44]
    What is the relationship between effective work function and electron ...
    Mar 17, 2015 · For example, a silicon crystal surface has electron affinity 4.05 eV, whereas an isolated silicon atom has electron affinity 1.39 eV. The ...
  45. [45]
    Ab initio calculation of ionization potential and electron affinity in ...
    Jan 22, 2016 · To accomplish that, we calculate the average electrostatic potential and vacuum level from slab models using the DFT method. The positions ...
  46. [46]
  47. [47]
    Trends on band alignments: Validity of Anderson's rule in S n S 2
    Apr 3, 2018 · According to this rule, the CBO in a semiconductor heterojunction has to be taken directly from the difference between the electron affinities, ...
  48. [48]
    General Properties of Silicon - PVEducation
    General Properties of Silicon ; Relative Permittivity (εr), 11.7 ; Electron Affinity, 4.05 eV ; Electron Diffusion Coefficient (De), kT/q µ ; Hole Diffusion ...
  49. [49]
    Latest advance on seamless metal-semiconductor contact with ...
    The interface feature is mainly governed by Schottky barrier (SB), which is defined as the difference between metal's work function and the electron affinity ( ...
  50. [50]
    GaN-Based Schottky Diode - IntechOpen
    Sep 12, 2018 · The difference between the semiconductor electron affinity and metal work function is defined as Schottky barrier height. The band structure ...
  51. [51]
    Schottky barriers on n-GaN grown on SiC | Journal of Electronic ...
    Characteristics of Schottky barriers fabricated on n-type GaN were investigated. The barriers were formed by vacuum thermal evaporation of Cr, Au, and Ni.
  52. [52]
    [PDF] Lecture 19 Metal-Semiconductors Interface
    metal and electron affinity of semiconductor, is called the Schottky barrier height, 𝜙SB = 𝜙m − 𝜒. The Schottky-Mott assumption is very handy, but often ...
  53. [53]
    Designing few-layer graphene Schottky contact solar cells
    In an ideal metal/semiconductor junction, the Schottky barrier height is dependent on the work function and the electron affinity of the semiconductor, as ...
  54. [54]
    Computational study of KGeCl3 perovskite solar cells toward high ...
    Sep 1, 2025 · The results reveal that for PCBM, an optimal electron affinity of 3.8 eV ensures ideal band alignment, promoting efficient charge transfer while ...
  55. [55]
    Solar Applications of Graphene - Cheap Tubes
    Electron transporter and acceptor based graphene-polymer solar cells rely on a high electron affinity to dissociate the electron-hole pairs into separate ...
  56. [56]
    Emerging Role of 2D Materials in Photovoltaics: Efficiency ...
    Aug 18, 2025 · Donor material efficiently absorbs photons and generates excitons while the acceptor material has a higher electron affinity than the donor ...
  57. [57]
    Elimination of charge-carrier trapping by molecular design - Nature
    Jun 29, 2023 · Trap-free bipolar charge transport can be accomplished when the electron affinity (EA) of the organic semiconductor is higher than 3.5 eV and ...
  58. [58]
    Single‐Layer Organic Light‐Emitting Diode with Trap‐Free Host ...
    Jan 12, 2024 · As good and balanced charge transport is a prerequisite for single-layer OLEDs, we first investigate the charge transport in pristine DMAC-BP.
  59. [59]
    Interface engineering for low-voltage operation of organic light ...
    Mar 25, 2025 · In this review, we introduce both the charge injection and recombination mechanisms in OLEDs and the interface control technology effective for lowering their ...
  60. [60]
    The Role of Electron Affinity in Determining Whether Fullerenes ...
    May 21, 2012 · Organic photovoltaics (OPVs) offer the potential of providing energy at a significantly lower cost than conventional inorganic technologies due ...
  61. [61]
    Tuning the electron affinity of cobalt oxide catalysts for robust acidic ...
    Jul 2, 2025 · The enhanced performance arises from the incorporation of Mo, whose higher electron affinity promotes *OH adsorption and steers the OER through ...
  62. [62]
    [PDF] Harnessing Electron Affinity to Enhance the Oxidative Performance ...
    Oct 24, 2025 · This study provides new evidence that electron affinity controls oxidation behavior in birnessite and offers a practical approach for designing ...