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).[1][2] 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.[2] In the periodic table, electron affinity generally becomes more negative (increases in magnitude) from left to right across a period, driven by increasing effective nuclear charge that pulls electrons closer to the nucleus despite similar atomic sizes.[3] It decreases down a group, as larger atomic radii and increased electron shielding reduce the attraction for additional electrons; halogens exhibit the highest affinities (e.g., fluorine at -328 kJ/mol), while alkali metals have lower values (e.g., lithium at -60 kJ/mol).[2] Exceptions occur, such as nitrogen's lower affinity than carbon due to electron repulsion in a half-filled p subshell, highlighting the role of electron configuration stability.[4] 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.[5] It also plays a key role in predicting lattice energies in solids and the behavior of elements in redox reactions, with high-affinity elements like halogens being strong oxidizing agents.[6] In advanced applications, such as semiconductor physics, electron affinity determines band structures and charge carrier mobility in materials like silicon.[7]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.[8] 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.[8] 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}.[2] The reverse process requires energy equal to the electron affinity value to detach the electron 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.[8]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.[9] A key interconnection arises in the Mulliken electronegativity scale, where electronegativity \chi is defined as the average of the ionization energy and electron affinity: \chi = \frac{\mathrm{IE} + \mathrm{EA}}{2} This formulation, proposed by Robert S. Mulliken 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.[10] 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.[11] 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.[12] 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 standard enthalpy change \Delta [H](/page/H+) under typical conditions (e.g., 298 K), where \Delta_\mathrm{eg} [H](/page/H+) \approx -\mathrm{EA} since thermal corrections are small and often negligible (~2-3 kJ/mol). The terms are sometimes used synonymously, particularly when thermal corrections are negligible, but \Delta_\mathrm{eg} [H](/page/H+) emphasizes the enthalpic aspect in thermochemical cycles like the Born-Haber process.[13]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 electron from a negative ion or the threshold 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 photoelectron spectroscopy, laser photodetachment, electron impact excitation, and equilibrium studies using ion traps or flow reactors.[14] 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.[14][15] 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[16] and 3.612720 ± 0.000044 eV for chlorine.[17] 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),[17] which is particularly advantageous for rare elements like actinides. Challenges include identifying the exact threshold amid weak signals or overlapping electronic states.[14] Electron impact methods involve colliding low-energy electrons with neutral atoms or molecules to form anions and measuring the threshold energy 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 function of incident electron energy, with thresholds extracted from onset curves. While historically important for early measurements (e.g., sulfur hexafluoride at 1.07 ± 0.07 eV), this approach suffers from lower precision, typically ±0.1-0.2 eV, due to background scattering 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.[14] Equilibrium studies derive electron affinities indirectly through thermochemical cycles involving gas-phase ion-molecule reactions, often measured in ion cyclotron resonance (ICR) mass spectrometers or flowing afterglow apparatuses. In ICR, equilibrium constants for reactions like A^- + B \rightleftharpoons A + B^- are determined from forward and reverse rate ratios at controlled temperatures, yielding \Delta H values related to EA(A) - EA(B). Flowing afterglow techniques similarly monitor ion intensities in a helium carrier gas to establish equilibria, as demonstrated for sulfur dioxide (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 temperature control and auxiliary bond dissociation energies.[14] Anion zero kinetic energy (ZEKE) spectroscopy represents a high-resolution extension of photodetachment methods, selectively detecting near-zero kinetic energy 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 lanthanide 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 noble gases and elements with near-zero affinities (e.g., uncertainties ~0.1 eV for helium 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.[14][18]Sign Convention and Units
In the early 20th century, the sign convention 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.[19] 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.[19] The modern convention, as defined by the IUPAC Gold Book, establishes electron affinity E_\text{ea} as the positive energy required to detach an electron from the singly charged negative ion in the gas phase, corresponding to the process \ce{X^- -> X + e^-}.[20] 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.[20] Older texts occasionally used \Delta E directly for attachment, resulting in negative values for exothermic cases, but the IUPAC approach aligns with ionization energy conventions for consistency.[20] 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}.[8] Values are often rounded for practical use; for example, the electron affinity of chlorine is 349 kJ/mol, equivalent to approximately 3.61 eV.[21] 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.[8] This relation facilitates connections between gas-phase ion data and formation enthalpies in thermochemical cycles.[8]Atomic Electron Affinities
Periodic Trends
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.[2] Down a group, electron affinity typically becomes less negative because the increasing atomic radius and electron shielding by inner shells weaken the nuclear pull on the incoming electron. In the halogen group (17), values become less negative from chlorine (-349 kJ/mol) to bromine (-325 kJ/mol) to iodine (-295 kJ/mol), with fluorine (-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 sulfur (-200 kJ/mol) but then less negative to selenium (-195 kJ/mol) and tellurium (-190 kJ/mol) due to size effects. Noble gases exhibit near-zero electron affinities, consistent with their filled valence shells that resist further electron addition.[2][22] Key factors influencing these trends include atomic radius, 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 nitrogen or noble gases) reduce affinity by increasing the energy required to accommodate an extra electron. These patterns are well-reproduced by computational methods such as density functional theory (DFT), which aligns predictions with experimental data for main group elements, often achieving mean absolute errors below 10 kJ/mol when using hybrid functionals.[2][23] 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:| Group | Element | Electron 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 halogens, where the electron affinity of fluorine (-328 kJ/mol) is less negative than that of chlorine (-349 kJ/mol), despite fluorine's higher effective nuclear charge. This anomaly stems from fluorine's exceptionally small atomic radius, which causes the added electron 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⁻.[2] Noble gases display particularly striking anomalies, with electron affinities that are generally near zero and often slightly positive (endothermic) due to their filled valence shells, making anion formation unfavorable in the ground state. For helium, the value is approximately 0 eV, though theoretical and experimental studies debate whether it is precisely zero or slightly positive (indicating an unbound ground state), with no stable anion observed. Heavier noble gases like xenon 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; ground states remain unbound.[24][25] In transition metals, electron affinities deviate irregularly from trends due to the variable stability of d-orbital configurations, often resulting in relatively low (less negative) values. For instance, chromium exhibits an electron affinity of -0.67 eV, attributed to the high stability of its half-filled 3d⁵ 4s¹ ground state, which requires electronic rearrangement upon adding an electron 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 effective nuclear charge and large size. Sodium, for example, has an electron affinity of -53 kJ/mol, indicating a small exothermic process for Na + e⁻ → Na⁻, but the anion is weakly bound due to the weak attraction for an additional electron 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 stability.[2] 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.[25]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.[26][27] 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.[28][29][26] 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 molecular geometry, often approximating the vertical detachment energy from the anion; in contrast, the adiabatic electron affinity accounts for structural relaxation in both the neutral and anion to their equilibrium geometries, which can differ substantially in polyatomic molecules due to vibrational and conformational adjustments.[26] 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.[26][30] 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.[31][26] 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.[27][32] 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 electron, 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 eV of experiment.[26] Density functional theory 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.[33][34]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 electron. For instance, benzene exhibits a negative electron affinity of approximately -1.15 eV, resulting in an unstable radical anion that rapidly loses the added electron due to the molecule's closed-shell aromatic structure.[35] In contrast, nitric oxide (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.[36] Halogen molecules like chlorine (Cl₂) show a more substantial positive value of 2.38 eV, facilitating stable anion formation due to the electronegative halogens.[37] 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.[28] 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 industrial contexts, electron affinity influences the design of dyes and conducting polymers, where high electron affinity promotes n-type doping and enhances charge transport. Studies in the 2020s have explored electron-deficient polymers with electron affinities around 2-3 eV to improve conductivity in organic electronics.[38] Recent research on molecular wires highlights how tuning electron affinity to values exceeding 2 eV enables efficient single-molecule conductance, as demonstrated in alkynylplatinum-based wires for nanoelectronic applications.[39] 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.| Molecule | Electron Affinity (eV) | Notes |
|---|---|---|
| O₂ | 0.451 ± 0.007 | Adiabatic value for stable atmospheric anion.[28] |
| NO | 0.026 ± 0.005 | Low value for odd-electron species.[36] |
| Cl₂ | 2.38 ± 0.10 | Positive, stable halogen anion.[37] |
| F₂ | 3.01 ± 0.07 | High value due to fluorine electronegativity.[40] |
| CO | 1.33 | Relevant for carbonyl reactivity.[41] |
| CO₂ | -0.60 | Negative, metastable radical anion.[42] |
| Benzene | -1.15 | Negative for aromatic stability.[35] |
| H₂O | ≈ -0.8 | Negative, diffuse anion state.[43] |