Fact-checked by Grok 2 weeks ago

18-electron rule

The 18-electron rule is a fundamental guideline in stating that stable complexes typically possess 18 electrons around the central metal atom, mimicking the electronic configuration and promoting thermodynamic . This rule, first articulated by in 1921 as an extension of theory to s, posits that the metal achieves a closed-shell structure by coordinating s that donate the requisite electrons to fill its , specifically the (n-1)d, ns, and np orbitals. Refined by Nevil Sidgwick in 1927 through the effective atomic number (EAN) concept, it emphasizes that the total electron count, derived from the metal's d electrons plus donations, correlates with observed in diamagnetic complexes. Electron counting under the 18-electron rule involves two common methods: the neutral ligand formalism, where ligands are treated as neutral donors (e.g., CO as a 2-electron donor), and the ionic formalism, which assigns formal charges to adjust the metal's before adding ligand electrons. For instance, in ([IrCl(CO)(PPh₃)₂]), the iridium center (group 9, d⁸ in oxidation state +1) receives 2 electrons from Cl, 2 from CO, and 2 each from the two phosphines, totaling 16 electrons, highlighting a common exception for square-planar d⁸ species. The rule applies primarily to low-spin, octahedral complexes of second- and third-row transition metals, where it rationalizes formulas like [Mn(CO)₆]⁺ (18 electrons) versus unstable alternatives. Beyond prediction, the 18-electron rule informs reaction mechanisms in , such as olefin and , where intermediates often alternate between 16-electron (unsaturated, reactive) and 18-electron (saturated, stable) configurations to facilitate steps like or ligand substitution. Exceptions abound, including 16-electron tetrahedral or square-planar complexes (e.g., Pd(0) catalysts), 14-electron species in early transition metals, and rare 20-electron cases stabilized by , as seen in a stable 20-electron derivative synthesized in July 2025 by Satoshi Takebayashi and colleagues at the Okinawa Institute of Science and Technology (OIST), challenging the rule's upper limit. These deviations underscore the rule's empirical , influenced by factors like metal , type, and , yet it remains a cornerstone for designing organometallic compounds in and .

Fundamentals

Electron Counting Methods

Electron counting in transition metal complexes is essential for applying the 18-electron rule, which posits that stable organometallic compounds typically achieve a total of 18 valence electrons around the central metal atom. Two equivalent conventions are widely used to perform this count: the neutral ligand method (also known as the covalent model) and the ionic ligand method (also known as the oxidation state formalism). These methods assign electron contributions from the metal and ligands in a manner that ensures consistency in the final tally, facilitating the prediction of complex stability without regard to bonding details. In the neutral ligand method, the metal is considered in its formal zero oxidation state, contributing a number of valence electrons equal to its group number in the periodic table (e.g., iron in group 8 contributes 8 electrons). Ligands are treated as neutral molecules or fragments, donating electrons based on their typical Lewis basicity: even-electron donors like carbon monoxide (CO) or phosphines (PR₃) contribute 2 electrons each, while odd-electron donors like halides (X) or hydrogen (H) contribute 1 electron each. The total electron count is the sum of the metal's contribution and all ligand donations, with adjustments for the overall charge of the complex if anionic or cationic (subtract or add electrons accordingly). This approach emphasizes covalent bonding character and is particularly straightforward for complexes with neutral ligands. The ionic ligand method, in contrast, begins by determining the of the metal, which is calculated by assigning full charges to s (e.g., halides as X⁻, cyclopentadienyl as Cp⁻) and balancing the complex's overall charge. The metal then contributes its d-electrons, given by the formula: d-electron count = group number - . s are classified as L-type (neutral, 2-electron donors like or PR₃) or X-type (anionic, 2-electron donors like Cl⁻ or H⁻ in their closed-shell form), with odd-electron ligands such as allyl often treated as 3-electron donors when neutral or 2-electron donors when anionic. The total is again the sum of metal d-electrons and ligand contributions, adjusted for charge; this method highlights ionic character and is useful for complexes with charged s. To perform the count using either method, the steps are as follows: (1) identify the metal and its relevant group number or ; (2) classify and tally donations from each ; (3) sum the metal and ; and (4) adjust for the complex's net charge if necessary (e.g., subtract 1 for a +1 charge). Common donations vary by method but are standardized for consistency:
LigandNeutral Method DonationIonic Method DonationNotes
2e (neutral σ-donor/π-acceptor)2e (L-type)Ubiquitous in carbonyl complexes.
PR₃ ()2e (neutral σ-donor)2e (L-type)Variable cone angle affects sterics.
(X, e.g., Cl)1e (neutral radical X•)2e (X-type, X⁻)Common in early complexes.
Allyl (C₃H₅)3e (neutral η³-allyl radical)2e (allyl⁻, X-type) or 4e (η³-allyl⁻ as L)Depends on and charge.
Cyclopentadienyl ()5e (neutral Cp•)6e (Cp⁻, L-type)Often η⁵-bound in compounds.
For illustration, consider pentacarbonyliron(0), Fe(CO)₅. Using the neutral method: iron (group 8) contributes 8 electrons, and each of the five CO ligands donates 2 electrons (total 10), yielding 18 electrons overall. The ionic method concurs: iron is in the 0 oxidation state (d⁸), with the same 10 electrons from five neutral CO ligands, for a total of 18 electrons. This example demonstrates the equivalence of the methods for homoleptic carbonyls.

Theoretical Basis

The 18-electron rule in chemistry arises from the configuration of d-block elements, which possess nine orbitals capable of accommodating up to 18 electrons for a stable, closed-shell arrangement. These orbitals include five (n-1)d, one ns, and three np orbitals, analogous to the in main-group elements where eight electrons fill the ns and np orbitals to achieve stability. For second- and third-row transition metals, this configuration mimics the electron setup of (Kr), with 18 outer electrons in a filled shell, providing thermodynamic stability through complete orbital occupancy. This principle was first proposed by in 1921 as part of his efforts to generalize the periodic table and to heavier elements, extending Lewis's ideas on chemical bonding to predict stable counts in metal compounds. In for octahedral complexes, the rule corresponds to filling six bonding molecular orbitals and three non-bonding t2g d-orbitals, resulting in no unpaired s in low-spin configurations. The ligand field plays a here, as strong π-acceptor ligands increase the octahedral splitting O), promoting low-spin filling that favors the 18- count for enhanced stability. The total valence electron count is formally expressed as the sum of the metal's d electrons and the electrons donated by ligands equaling 18 for optimal stability: \text{Total electrons} = n + \sum \text{(ligand electrons)} = 18 where n represents the number of valence d electrons from the metal center. This equation underscores the rule's basis in achieving a filled valence shell, though it serves as a guideline rather than a strict law due to variations in coordination geometry and ligand interactions.

Applicability

Suitable Metal Centers and Ligands

The 18-electron rule is most reliably applicable to middle and , particularly those in groups 6 through 10 of the periodic table, such as the , , iron, triads. These metals, when in low oxidation states (typically +0 to +2), can accommodate sufficient ligands to achieve the 18-electron configuration, mimicking the stability of noble gas electron shells. In , early transition metals in groups 3 through 5 often fail to follow the rule due to their higher and larger ionic radii, which limit ligand coordination and result in electron counts below 18, even in stable complexes. Essential for adherence to the 18-electron rule are π-acceptor ligands, which stabilize low-oxidation-state metals through synergistic σ-donation and π-backbonding from the metal's filled d-orbitals./Advanced_Inorganic_Chemistry_(Wikibook)/01%3A_Chapters/1.19%3A_Electron_Counting_and_the_18_Electron_Rule) Representative examples include carbon monoxide (CO), nitrosyl cation (NO⁺), and phosphines (PR₃), which effectively delocalize electron density from the metal center, facilitating the required electron count./Advanced_Inorganic_Chemistry_(Wikibook)/01%3A_Chapters/1.19%3A_Electron_Counting_and_the_18_Electron_Rule) Ligands lacking significant π-acceptor ability, such as certain soft σ-donors without backbonding capacity, render the rule less predictive, as they do not sufficiently stabilize the electron-rich metal environment./Advanced_Inorganic_Chemistry_(Wikibook)/01%3A_Chapters/1.19%3A_Electron_Counting_and_the_18_Electron_Rule) The applicability of the 18-electron rule further depends on low-spin electron configurations at the metal center, which are promoted by strong ligand fields that minimize unpaired electrons./4%3A_Crystal_Field_Theory/4.3%3A_High_Spin_and_Low_Spin_Complexes) This requires the ligand field splitting parameter (Δ) to exceed the pairing energy (P), ensuring electrons occupy lower-energy orbitals preferentially./4%3A_Crystal_Field_Theory/4.3%3A_High_Spin_and_Low_Spin_Complexes) Weak-field ligands, where Δ < P, lead to high-spin states that disrupt the closed-shell stability predicted by the rule./4%3A_Crystal_Field_Theory/4.3%3A_High_Spin_and_Low_Spin_Complexes)

Classic Examples

One of the most iconic examples of the 18-electron rule is ferrocene, [Fe(η⁵-C₅H₅)₂], where the iron(II) center (d⁶ configuration) receives 6 electrons from the metal and 6 electrons each from the two cyclopentadienyl (Cp⁻) ligands, yielding a total of 18 valence electrons. This complex adopts a sandwich structure with parallel Cp rings, which correlates with its achievement of the 18-electron count and contributes to its exceptional stability, allowing isolation as an air-stable, crystalline solid. Iron pentacarbonyl, Fe(CO)₅, exemplifies the rule for mononuclear carbonyl complexes, with the iron(0) center (d⁸) donating 8 electrons and each of the five CO ligands contributing 2 electrons via σ-donation and π-backbonding, for a total of 18 electrons. The molecule exhibits trigonal bipyramidal geometry, a structural adjustment that accommodates the five ligands to satisfy the electron count, and it is readily synthesized and isolated under mild conditions due to this saturated configuration. Chromium hexacarbonyl, Cr(CO)₆, represents a homoleptic octahedral complex adhering to the 18-electron rule, where the chromium(0) d⁶ center provides 6 electrons supplemented by 12 from the six CO ligands (2 electrons each). This geometry is typical for d⁶ metals achieving 18 electrons, enabling the complex to be prepared via direct reaction of chromium metal with CO and isolated as a colorless, volatile solid that remains unreactive toward many reagents. The cationic complex [Mn(CO)₆]⁺ illustrates the rule's applicability to ionic species, with the manganese(I) d⁶ center contributing 6 electrons and the six CO ligands adding 12, totaling 18 electrons in an octahedral arrangement./24%3A_Organometallic_chemistry-_d-block_elements/24.03%3A_The_18-electron_Rule) Such complexes are synthetically accessible through oxidation of neutral precursors and are isolable as stable salts, underscoring how the 18-electron configuration predicts robust structures across diverse coordination environments./24%3A_Organometallic_chemistry-_d-block_elements/24.03%3A_The_18-electron_Rule)

Stability and Reactivity

Stability Consequences

Low-spin 18-electron complexes typically exhibit a closed-shell electronic configuration with no unpaired electrons, resulting in diamagnetic properties and inherently lower reactivity compared to open-shell species. This configuration corresponds to a filled valence shell of ns²(n-1)d¹⁰np⁶, analogous to noble gas stability, which minimizes the availability of electrons for redox or radical processes. The stability of these complexes is further enhanced by robust metal-ligand bonds, arising from the synergistic interplay of σ-donation from ligands to the metal and π-backbonding from the metal d-orbitals to ligand π* orbitals. This mutual electron sharing strengthens the overall bonding framework, particularly with π-acceptor ligands like CO, which effectively delocalize electron density and prevent destabilizing accumulation on the metal center. Thermodynamically, 18-electron species demonstrate greater stability than their unsaturated counterparts, as evidenced by higher formation constants that reflect favorable equilibrium positions for their assembly. For instance, the large ligand field splitting (Δ_O) in these complexes avoids occupation of antibonding e_g orbitals, lowering the overall energy and promoting persistence under ambient conditions. Kinetically, 18-electron complexes often display exchange-inert behavior due to the high activation barriers for ligand substitution, which require dissociation to a less stable 16-electron intermediate. A representative example is , Cr(CO)_6, where ligand exchange proceeds via a dissociative mechanism at elevated temperatures, with rate constants on the order of 10^{-5} s^{-1} at 130°C, underscoring its remarkable inertness relative to labile, electron-deficient analogs. This 18-electron paradigm parallels the octet rule in main-group chemistry, where both represent closed-shell configurations that confer stability through complete valence orbital filling; however, the transition metal variant accommodates delocalized d-orbital bonding, enabling diverse coordination geometries absent in p-block elements.

Reactivity Patterns

The 18-electron configuration imparts significant inertness to organometallic complexes, particularly with respect to associative ligand substitution pathways. In such species, the filled valence shell discourages direct addition of incoming ligands without prior dissociation, making associative mechanisms rare and favoring slower dissociative routes where a ligand departs first to generate a coordinatively unsaturated intermediate. For instance, in Ni(CO)4, an 18-electron complex, CO exchange proceeds via a dissociative mechanism, as associative attack is energetically prohibitive. This kinetic stability arises from the electronic saturation, which minimizes the driving force for bond formation without relief of steric or electronic crowding. In contrast, 16-electron precursors serve as highly reactive species in many catalytic processes, often acting as initiators that readily coordinate substrates to achieve the 18-electron state. A prominent example is found in olefin metathesis catalysis, where precatalysts like the first-generation Grubbs complex, [RuCl2(=CHPh)(PCy3)2], a 16-electron species, dissociate a ligand to form a 14-electron active center that coordinates , initiating the metathesis cycle through metallacyclobutane formation. This unsaturation enables facile substrate binding and turnover, highlighting how deviations from 18 electrons drive reactivity in unsaturated complexes. Oxidative addition reactions exemplify how 16-electron metals achieve the 18-electron configuration by incorporating substrates like dihydrogen or alkyl halides. The classic case is , IrCl(CO)(PPh3)2, a 16-electron species that undergoes rapid of H2 to form the 18-electron dihydride IrH2Cl(CO)(PPh3)2, increasing the and electron count simultaneously. This step is a cornerstone of many catalytic cycles, as the resulting 18-electron intermediate provides a stable platform for subsequent transformations. Conversely, serves as the reverse process, converting 18-electron complexes back to 16-electron species by extruding a , such as in the elimination of H2 from cis-IrH2Cl(CO)(PPh3)2 to regenerate the active 16-electron catalyst. This elimination is often rate-limiting in cycles like , underscoring the energetic barrier to relieving the 18-electron saturation. Historically, the 18-electron rule, formalized in the , provided a unifying framework for understanding why numerous homogeneous catalytic processes, such as and Wilkinson's , proceed through alternating 16-electron and 18-electron states rather than remaining locked in one configuration. This alternation accommodates the need for substrate binding and product release, explaining the prevalence of in otherwise stable 18-electron precursors and the associative reactivity of unsaturated intermediates. The rule's predictive power for reaction mechanisms has enduringly shaped organometallic design, emphasizing pathways that preserve effective counts during .

Exceptions

16-Electron Complexes

16-electron complexes constitute a prominent exception to the 18-electron rule, particularly among d⁸ transition metals that adopt a square-planar , resulting in an effective d¹⁰ electronic configuration analogous to a closed-shell setup. These are coordinatively unsaturated, featuring an empty orbital in the metal's valence shell, which contrasts with the fully saturated 18-electron octahedral complexes typical of early transition metals. For late transition metals such as those in groups 8–10, this 16-electron count is viable due to the higher energy of antibonding orbitals in square-planar arrangements, minimizing electronic repulsion. The stability of these 16-electron complexes often arises from steric saturation provided by bulky ligands, which prevent additional coordination, or from the inherent tolerance of late metals to coordinative unsaturation without destabilization. In d⁸ systems like Rh(I) and Ir(I), the square-planar geometry is preferred because it allows strong σ-donation from four ligands while avoiding the pairing energy penalties of octahedral filling. Examples include , \ce{RhCl(PPh3)3}, a 16-electron square-planar complex active in , where the ligands provide steric bulk to maintain the unsaturated state. Similarly, , \ce{IrCl(CO)(PPh3)2}, is a stable 16-electron species that undergoes reactions, highlighting its electronic unsaturation. Geometrically, these complexes typically exhibit square-planar coordination with bond angles near 90°, though some three-coordinate variants adopt T-shaped structures to accommodate the 16-electron count while leaving a site vacant for reactivity. In catalysis, the unsaturated of 16-electron complexes enables direct binding without the need for dissociation, facilitating steps like in processes such as . This feature underscores their role as precatalysts in homogeneous systems, where the empty orbital promotes associative mechanisms.

Ligand and Spin State Effects

The presence of bulky introduces steric hindrance around the metal , which can prevent the coordination of additional ligands necessary to reach the 18-electron configuration, favoring lower electron counts for enhanced stability. In late transition metals, triphenylphosphine (PPh₃) exemplifies this effect; the tetrahedral Ni(PPh₃)₄ complex nominally achieves 18 electrons but is unstable due to crowding from the ligands' large cone (approximately °), leading to rapid dissociation in to the trigonal planar 16-electron Ni(PPh₃)₃ species and free PPh₃. Similarly, substitution with even bulkier phosphines, such as P(t-Bu)₃, further enforces low coordination numbers in (0) systems, stabilizing unsaturated complexes that resist additional . π-Donating ligands, particularly strong donors like or , can stabilize es with counts below 18 by providing excess through π-donation from p-orbitals to metal d-orbitals, effectively reducing the demand for more ligands while maintaining coordinative saturation. The octahedral [TiF₆]²⁻ anion illustrates this for early transition metals: Ti(IV) contributes 0 d-s, and the six F⁻ ligands donate 12 s, yielding a 12- where fluoride's π-donation lowers the octahedral splitting by mixing with t_{2g} orbitals, stabilizing the low count by keeping antibonding character unoccupied in d⁰ systems without bond weakening. This effect is pronounced in d⁰ systems, where π-donors enable high coordination numbers despite sub-18 configurations. High-spin configurations arise with weak-field ligands that induce minimal crystal field splitting, resulting in unpaired d-electrons and an effective deviation from the 18-electron rule, as the guideline assumes low-spin, diamagnetic species with filled t₂g orbitals. The hexaaquamanganese(II) ion [Mn(H₂O)₆]²⁺ serves as a key example: Mn²⁺ (d⁵) provides 5 electrons, and the six aqua ligands contribute 12, for a total of 17 valence electrons in a high-spin (S = 5/2) state with five unpaired electrons, yet the complex remains stable in due to the character and hydration shell. Such high-spin cases are common in first-row transition metals with neutral or weakly donating ligands like or . Combinations of bulky and π-donating ligands amplify these exceptions by simultaneously imposing steric constraints and electronic stabilization, often through multidentate designs that rigidify and tune . For instance, chelating diphosphines with π-donating substituents in complexes enforce square-planar 16-electron geometries, where bulk prevents axial while π-donation supports the unsaturated state, enhancing reactivity in . Spectroscopic evidence, including (EPR) and magnetic measurements, distinguishes these ligand- and spin-induced deviations by detecting unpaired s and confirming high-spin ground states. In high-spin Mn(II) aqua complexes, EPR spectra exhibit isotropic signals at g ≈ 2.0 with hyperfine splitting from the ⁵⁵Mn (I = 5/2), verifying the S = 5/2 configuration and the associated electron count irregularity.

Higher Electron Counts

Complexes exceeding the 18-electron count are relatively rare in transition metal chemistry, as additional electrons typically occupy antibonding orbitals, leading to weakened metal-ligand bonds. However, certain structural and electronic factors can stabilize such species, including anionic charges that mitigate electrostatic repulsion and chelating ligands that impose geometric constraints to prevent dissociation. For instance, the 19-electron sandwich compound cobaltocene, [Co(η⁵-C₅H₅)₂], is a well-known paramagnetic example where the odd electron resides in a metal-based orbital, rendering it air-sensitive but isolable under inert conditions. Similarly, 20-electron octahedral carbonyl anions like [Sc(CO)₈]⁻, [Y(CO)₈]⁻, and [La(CO)₈]⁻ have been synthesized and characterized in the gas phase, demonstrating that group 3 metals can support higher counts through enhanced back-donation into CO π* orbitals. Early metals in groups 3–5 particularly tolerate counts above 18 due to their larger radii and more diffuse d-orbitals, which accommodate additional ligands without severe steric or electronic penalty. A representative case is the 20- hexamethyl tantalate anion, [Ta(CH₃)₆]⁻, which adopts a trigonal prismatic and was prepared via of TaCl₅ followed by , highlighting the role of alkyl ligands in providing high . These larger orbitals facilitate better overlap with donor ligands, allowing population of otherwise destabilizing levels while maintaining overall complex integrity. Despite these stabilizing features, higher electron counts generally confer reduced thermodynamic stability compared to 18-electron analogs, often manifesting as a propensity for ligand elimination to relieve antibonding interactions. For example, 19- and 20-electron species like cobaltocene derivatives readily lose ligands or undergo reactions under mild conditions to approach the octet-like closure. Recent computational studies have further elucidated this, showing that in cases like [TM(CO)₈]⁻ (TM = group 3), the extra electrons occupy delocalized bonding orbitals with minimal disruption to the core structure, confirming viability for specific late main-group mimics but underscoring inherent lability in solution.

Duodectet Rule

The duodectet rule describes a stable of 12 electrons for main-group elements and early metals, particularly in tetrahedral or polyhedral cluster geometries, where bonding involves six orbitals consisting of the s orbital and five d-like orbitals, excluding p-orbital participation. This configuration achieves a closed-shell arrangement through sd-hybridization and three-center/four-electron (3c/4e) resonance bonding, providing enhanced stability without invoking hypervalency via p orbitals. Computational analyses using (NBO) methods support this model, showing that the p orbitals remain largely uninvolved and empty in such systems. The rule is particularly applicable to boranes, carboranes, and early clusters, where the skeletal electron count of 12 electrons corresponds to stable closo structures as predicted by Wade's rules for polyhedral . For instance, the [B₅H₅]²⁻ anion features 12 skeletal electrons (six pairs) in a trigonal bipyramidal , with each boron-hydrogen unit contributing to delocalized cluster bonding via 3c/2e interactions for the skeletal framework. Similarly, carboranes like C₂B₃H₅ follow analogous counting, maintaining 12 skeletal electrons for closo-tetrahedral derivatives. Early clusters, such as those involving or with ligands, also exhibit stability at this electron count through comparable delocalized bonding. Both the duodectet and 18-electron rules stem from of closed-shell polyhedral electron counts, with Wade's rules serving as a foundational extension that rationalizes geometries based on skeletal electron pairs (n+1 pairs for n vertices in closo-boranes). However, the duodectet rule differs by focusing on bonding in main-group systems, emphasizing multicenter delocalized interactions rather than the ligand-filled coordination spheres of mononuclear complexes (MLₓ) under the 18-electron paradigm. A representative example is the neutral Al₄ , which adopts a tetrahedral in the gas with its 12 electrons (three per aluminum atom) filling the bonding molecular orbitals, conferring exceptional stability relative to other isomers.

Modern Applications and Insights

In contemporary catalysis, the 18-electron rule informs the design of cycles where species alternate between 16- and 18-electron states to promote key steps like substrate coordination and insertion. For olefin polymerization with Ziegler-Natta catalysts, early transition metal centers such as titanium operate in high oxidation states (e.g., Ti(III) or Ti(IV)), often with electron counts below 18 due to coordinative unsaturation, which allows for the coordination of monomers like ethylene in the catalytic cycle, achieving high stereoselectivity in polymer formation. In palladium-catalyzed cross-coupling reactions, such as Suzuki-Miyaura couplings, the mechanism proceeds through 14- to 16-electron intermediates, including monoligated Pd(0) species that deviate from the 18-electron saturation to facilitate of aryl halides. These unsaturated states enhance reactivity by lowering the energy barrier for substrate binding, with computational modeling confirming that such low-ligated Pd centers outperform traditional 18-electron precatalysts in turnover frequency. Density functional theory (DFT) studies since 2019 have refined insights into distributions, showing that the 18-electron rule arises from optimized orbital filling and effects in complexes, rather than rigid valence shell closure. For example, post-2019 DFT analyses of d-block systems reveal how electron delocalization influences stability, with deviations linked to ligand field strengths and states. In the 2020s, extensions to f-block analogs, such as and organometallics, demonstrate the rule's limitations, as non-bonding f-orbitals permit stable 14- to 20-electron configurations without the bonding constraints seen in d-block elements. Notable recent applications include 2022 DFT investigations of 18-electron ruthenium(II) complexes for transfer hydrogenation of ketones, where the precatalyst activates to a 16-electron hydride species that delivers H⁻ efficiently using isopropanol as the hydrogen donor, achieving high enantioselectivity in asymmetric variants. In bioinorganic contexts, iron-sulfur (Fe-S) clusters in enzymes like [FeFe]-hydrogenases deviate from the 18-electron rule through delocalized electrons across the cluster, enabling redox flexibility for proton reduction, as confirmed by DFT modeling of their H-cluster states. Looking ahead, the 18-electron rule is being adapted to and single-atom catalysts (SACs), where metal centers anchored on supports like TiO₂ mimic coordination saturation; for instance, SACs reaching 18 electrons via support interactions suppress unwanted adsorption, enhancing selectivity in oxidation reactions. A 2019 Chemical Reviews article on computational approaches to 3d-metal underscores electron counting's role in predicting active sites and optimizing environments for sustainable processes.

References

  1. [1]
    Types of Valence - Science
    Types of Valence. Irving LangmuirAuthors Info & Affiliations. Science. 22 Jul ... PDF format. Download this article as a PDF file. Download PDF. Figures. Tables ...
  2. [2]
    The Electronic Theory Of Valency : Nevil Vincent Sidgwick
    Oct 26, 2006 · The Electronic Theory Of Valency. by: Nevil Vincent Sidgwick. Publication date: 1927. Topics: Chemistry. Publisher: Oxford At The Clarendon ...
  3. [3]
    The 16 and 18 electron rule in organometallic chemistry and ...
    The 16 and 18 electron rule in organometallic chemistry and homogeneous catalysis. CA Tolman, Chem. Soc. Rev., 1972, 1, 337.
  4. [4]
    Organometallic compound found to break textbook principle
    Jul 15, 2025 · A new derivative of ferrocene has 20 valence electrons, challenging the longstanding rule that organometallic compounds can only have a maximum of 18.Missing: primary | Show results with:primary
  5. [5]
    The 18-electron rule and electron counting in transition metal ...
    Mar 5, 2015 · An alternative electron count model, known as the effective atomic number (EAN) rule, was introduced in 1927 by Nevil Sidgwick (1873–1952) [5] ...
  6. [6]
    The Origin of the 18-Electron Rule | Journal of Chemical Education
    In response to a reader query, this article traces the origins of the 18-electron rule back to a 1921 paper by Irving Langmuir.
  7. [7]
    [PDF] Platinum (Pt ) , a 4d transition metal, which forms a large number of ...
    Middle transition metals (group 6 to group 8) commonly have 18 electron count while late transition metals (group 9 to group 11) generally have 16 or lower ...<|separator|>
  8. [8]
    The Electronic Structure of Ferrocene - AIP Publishing
    The 18 valence electrons of the compound are distributed among these MO's so that 16 occupy bonding, and 2 nonbonding MO's. This electron distribution ...Missing: paper | Show results with:paper
  9. [9]
    [PDF] The 16 and 18 Electron Rule in Organometallic Chemistry and ...
    The accessibility of 16 and 18 electron configurations also has important consequences for mechanisms of organometallic reactions, as outlined below.<|control11|><|separator|>
  10. [10]
    Nineteen-electron complexes and their role in organometallic ...
    ... Electron Transition Metal Complex. The Journal of Physical Chemistry A ... Angewandte Chemie 2018, 130 (30) , 9563-9568. https://doi.org/10.1002/ange ...
  11. [11]
    Valence and extra‐valence orbitals in main group and transition ...
    Oct 24, 2006 · ... duodectet rule” in the NBO Lewis-like description. Let us first illustrate use of the RMSD criterion to compare single-reference vs. multi ...
  12. [12]
    Journal of the Chemical Society D - RSC Publishing
    K. Wade. Abstract. The skeletal structures of carboranes, the higher boranes and borane anions, and many transition-metal carbonyl cluster species are related ...
  13. [13]
    Theoretical Study of the O2 Interaction with a Tetrahedral Al4 Cluster
    The bare tetrahedral structure is shown to be ∼10 kcal/mol above the ground state of the bare rhombus. Having ascertained the energy of the optimized ...Missing: structure Jones
  14. [14]
    Olefin Polymerization with Ziegler-Natta Catalyst
    Jun 30, 2023 · The Ziegler-Natta (ZN) catalyst, named after two chemists: Karl Ziegler and Giulio Natta, is a powerful tool to polymerize α-olefins with high linearity and ...
  15. [15]
    Recent developments in redox-active olefin polymerization catalysts
    Oct 1, 2018 · This review discusses the recent evolution of redox-active olefin polymerization catalysts, which has extensively focused on the use of ...
  16. [16]
    Emerging Trends in Cross-Coupling: Twelve-Electron-Based L1Pd ...
    Oct 3, 2022 · ... 18-electron rule postulated by Langmuir. (35,36). Figure 1. Figure 1 ... Novel electron-rich bulky phosphine ligands facilitate the palladium- ...
  17. [17]
    Revealing the nature of electron correlation in transition metal ...
    May 20, 2021 · In this work, we provide a nuanced view of electron correlation in the context of transition metal complexes, reconciling computational ...<|separator|>
  18. [18]
    A Density Functional Theory for the Average Electron Energy
    Jan 24, 2023 · A formally exact density functional theory (DFT) determination of the average electron energy is presented.
  19. [19]
    Actinide-lanthanide single electron metal-metal bond formed in ...
    Oct 20, 2023 · Understanding metal-metal bonding involving f-block elements has been challenging. Here, the authors report a series of mixed-valence di- ...
  20. [20]
    Density functional theory investigation of Ru(II) and Os(II ...
    Feb 14, 2022 · When the catalyst is chiral (natural or synthetic), transfer hydrogenation is often enantioselective (asymmetric transfer hydrogenation, ATH).
  21. [21]
    Enzymatic Mechanism of Fe-Only Hydrogenase: Density Functional ...
    In [FeFe]H, both of the iron ions tend to satisfy the 18-electron rule. Thus ... Fe-S clusters instead of Ni in their active sites. 27. Adams, M. W. W. ...Abstract · Computational Details · Supporting Information...
  22. [22]
    Strong Metal-Support Interactions between Pt Single Atoms and TiO2
    Jul 13, 2020 · The suppression of CO adsorption on Pt single atoms stems from coordination saturation (18-electron rule) rather than the physical coverage of ...Missing: nanomaterials | Show results with:nanomaterials
  23. [23]
    Nickel complexes as molecular catalysts for water splitting and CO2 ...
    Jan 1, 2019 · 2019, Chemical Reviews. Computational Approach to Molecular Catalysis by 3d Transition Metals: Challenges and Opportunities. 2019, Chemical ...
  24. [24]
    From 18- to 20-electron ferrocene derivatives via ligand coordination
    Peer-reviewed article in Nature Communications describing the synthesis and stability of 20-electron ferrocene derivatives by Satoshi Takebayashi and colleagues at OIST, published July 7, 2025.