Fact-checked by Grok 2 weeks ago

Reductive elimination

Reductive elimination is a key in in which two cis-disposed ligands on a center couple to form a new σ-bond between them, concomitant with a two-electron reduction of the metal's formal and a decrease in its . This process, the microscopic reverse of , typically involves the formation of C–C, C–H, C–X (where X is a or other ), or related bonds, and is thermodynamically driven by the stability of the newly formed bond relative to the metal–ligand bonds. It proceeds via a concerted two-electron in most cases, though single-electron pathways involving intermediates have been observed in specific systems, such as certain - or palladium-catalyzed processes. As a cornerstone of catalysis, reductive elimination serves as the product-releasing step in numerous cross-coupling reactions, including the Suzuki–Miyaura, Heck, and Negishi couplings, where it facilitates the construction of carbon–carbon and carbon–heteroatom bonds essential for . The reaction's efficiency is highly influenced by factors such as the metal's , steric hindrance from supporting ligands (e.g., phosphines or N-heterocyclic carbenes), and the nature of the ligands involved; for instance, β-hydrogen elimination often competes with reductive elimination in alkyl metal complexes, leading to alternative decomposition pathways. While prevalent in late s like , , and , reductive elimination has also been documented in early metals and f-block elements, such as complexes, expanding its scope beyond traditional d-block catalysis. Recent advances have leveraged reductive elimination in innovative contexts, including oxidatively induced variants that enable selective bond formations under mild conditions. These developments underscore its versatility, with ongoing research focusing on modulating environments to overcome kinetic barriers and enhance selectivity in complex synthetic transformations.

Fundamentals

Definition and Scope

Reductive elimination is a fundamental elementary step in , defined as a two-electron process in which two cis-disposed ligands bound to a center form a new between them and depart from the complex, thereby decreasing the formal of the metal by two units. This process simultaneously reduces the of the metal by two and increases its d-electron count by two. The general reaction can be schematically represented as \ce{ML2 ->[reductive elimination] M + L-L}, where \ce{M} denotes the metal fragment and \ce{L} represents the departing ligands, which may include alkyl, aryl, , or other groups capable of forming stable bonds such as C–C, C–H, or C–X. Unlike β- elimination, which involves the intramolecular transfer of a from a β-position to the metal with concomitant formation, reductive elimination requires direct coupling of two distinct ligands without such β- involvement. This transformation is particularly prevalent in complexes with d⁸ to d¹⁰ electron configurations, such as those of Pd(II), Pt(II), Ni(II), and Au(III), where it serves as a critical step for product release in numerous catalytic cycles, including cross-coupling reactions and C–H functionalizations. Reductive elimination is the microscopic reverse of , enabling reversible bond activation and formation in organometallic transformations. Thermodynamically, reductive elimination is often exergonic, driven by the formation of a strong L–L bond that compensates for the cleavage of two relatively weaker M–L bonds, especially in higher oxidation states where metal–ligand interactions are diminished. This energetic favorability underpins its role as a driving force in many catalytic processes, though the kinetics can vary depending on the specific ligands and metal.

Historical Development

The concept of reductive elimination emerged in the early through observations in - and -catalyzed reactions. These early studies highlighted the process in stoichiometric contexts, where dialkylmetal complexes decomposed to form alkanes, but the mechanistic role remained unclear until later investigations. Formal recognition of reductive elimination as a distinct elementary step occurred in the , with David Milstein and John K. Stille providing key evidence through their studies on complexes in the Stille . In 1978, they demonstrated that from organotin reagents to alkylpalladium(II) species precedes cis and subsequent reductive elimination to coupled products, establishing the process as central to cross-coupling mechanisms. A pivotal milestone came in 1974 with ' report on alkyl group migrations in and systems, where stereospecific reductive elimination from cis-dialkyl complexes formed carbon-carbon bonds, offering direct insight into migratory pathways. During the , advancements in , including the development of the Heck and reactions, further linked reductive elimination to efficient catalytic cycles, shifting focus from isolated stoichiometric events to productive turnover in . The understanding of reductive elimination evolved significantly from stoichiometric demonstrations to integral components of catalytic processes by the late . In the , computational studies have confirmed and refined these pathways, revealing details of transition states and electronic factors; for instance, calculations have elucidated two-electron processes in and systems, supporting experimental observations of barriers and influences. Later contributions by John F. Hartwig integrated reductive elimination into C-H strategies, demonstrating its utility in forming carbon-heteroatom bonds from high-valent intermediates in - and -catalyzed functionalizations.

Reaction Mechanisms

Octahedral Complexes

Reductive elimination in octahedral metal complexes, typically those of d⁶–d⁸ configurations such as (IV) and (III), requires the eliminating ligands to occupy positions within the . Trans arrangements inhibit the process due to insufficient orbital overlap between the ligand-based σ orbitals, preventing effective . This geometric constraint ensures proximity for bond formation, as seen in () isomers of octahedral (IV) species where adjacent methyl groups facilitate elimination. The mechanism proceeds through distinct steps: initial ligand migration involving a three-center transition state that aligns the cis ligands for coupling, followed by formation of the new C–C or C–H bond, and subsequent release of the product to yield a lower-valent metal species. In many cases, especially for C–C bond formation, a prior dissociation of a ligand (such as a phosphine) generates a five-coordinate intermediate, enabling the migratory step; C–H elimination often occurs directly from the saturated octahedral center. Product release completes the cycle, reducing the coordination number and oxidation state by two units. This process requires thermal activation at moderate temperatures, consistent with experimental observations. A representative example is the C–C reductive elimination from fac-(dppe)PtMe₄, which forms ethane and (dppe)PtMe₂: \text{fac-(dppe)PtMe}_4 \rightarrow \text{(dppe)PtMe}_2 + \text{C}_2\text{H}_6 Orbital analysis reveals that d orbitals, particularly the d_{z^2} component, play a crucial role in facilitating σ bond formation by mediating electron transfer from metal–ligand bonds to the nascent ligand–ligand interaction. In Pt(IV) systems, the high-lying metal-based HOMO incorporates d character to overlap with ligand σ* orbitals in the transition state, promoting coupling. Similar involvement occurs in Ir(III) complexes, where cis alkyl–hydride elimination benefits from compatible d–σ hybridization for efficient bond formation.

Square Planar Complexes

Reductive elimination from square planar d⁸ complexes, such as those of Pd(II) and Ni(II), is promoted by the inherent geometry, which positions cis-oriented ligands in proximity for while disfavoring trans elimination due to the 90° bond angles. This process commonly occurs in 16-electron species, where the planar arrangement enables efficient overlap in the without the steric congestion typical of higher-coordinate systems. The mechanism generally proceeds in three key steps: if the eliminating groups occupy trans positions, isomerization to the cis configuration occurs first, often via phosphine dissociation and recoordination; this is followed by concerted migration of the groups through a four-center involving partial bond breaking and formation; finally, the dissociates rapidly, yielding a 14-electron metal(0) species. Kinetic studies reveal dependence on the complex concentration, consistent with a rate-determining step at or before the . A prototypical example is the cis elimination from Pd(PEt₃)₂(Ph)(Me) to form Pd(PEt₃)₂ and (PhMe), proceeding via the described pathway with unimolecular (rate = k[complex]). This reaction is typically 10³–10⁶ times faster than analogous eliminations from octahedral complexes, attributable to the accessible low-coordination intermediates and reduced entropic penalties in the square planar geometry. Experimental support includes ¹H NMR crossover experiments, such as those with isotopically labeled dimethylpalladium analogs, which yield only homodimers (e.g., and perdeuteroethane) and confirm the intramolecular, cis-selective nature without intermolecular exchange. Computational DFT analyses, using functionals like MPWB1K, model the four-center with low barriers of approximately 15 kcal/, aligning with the observed rapid s at mild temperatures. Variations arise in 14-electron precursors bearing a vacant coordination site, where elimination can occur directly from the three-coordinate species without prior dissociation, further accelerating the process in coordinatively unsaturated systems.

Influencing Factors

Electronic and Metal Effects

Late transition metals such as , , and generally facilitate reductive elimination more readily than early transition metals like , owing to weaker metal-carbon bonds that lower the energetic barrier for bond cleavage and formation. Early metals resist this process due to their lower , higher oxidation states, and stronger M-C interactions, which stabilize the organometallic intermediates. Among group 10 metals, the rate of elimination follows the trend Ni > Pd > Pt, reflecting differences in orbital overlap and bond strengths that influence the stability. The electron density at the metal center plays a crucial role in reductive elimination, with electron-poor metals promoting the process by facilitating the transfer of electron density from the metal to the departing ligands in the transition state. Conversely, electron-rich metals, often resulting from strong π-backbonding with ancillary ligands, stabilize higher oxidation states and thereby slow reductive elimination; for instance, platinum(II) complexes exhibit slower rates compared to nickel(0) precursors in certain cycles due to enhanced backbonding that populates metal d-orbitals. This effect is particularly pronounced in the reduction from higher oxidation states, such as M(IV) to M(II), where decreased electron density accelerates the elimination to achieve a more stable lower-valent product. Quantitative trends underscore these electronic influences, as demonstrated by Hammett studies on aryl-substituted complexes, which reveal negative rho values (e.g., ρ = -1.36 for C-O elimination in complexes), indicating that electron-withdrawing substituents on aryl ligands accelerate reductive elimination by destabilizing the metal-aryl bond. These studies highlight a with metal d- count, where higher d^8 configurations in late metals like and Pt(II) exhibit modulated rates based on electron withdrawal, with fewer d-electrons facilitating faster elimination. Tolman electronic parameters (TEP) for ligands further link ancillary ligand electronics to metal density, showing that more electron-donating ligands (lower TEP) indirectly influence elimination rates by enhancing backbonding, though this intersects with steric factors.

Steric and Ligand Effects

Steric effects significantly influence the rate of reductive elimination in complexes by destabilizing the relative to the , particularly in square planar d8 systems like (II). Bulky ancillary increase steric congestion around the metal center, which alleviates upon elimination of the product and thereby accelerates the process. For instance, tri-tert-butylphosphine (P(t-Bu)3), with its large cone angle, promotes faster reductive elimination in crowded complexes compared to less bulky phosphines like PPh3, as the strain in the starting complex is relieved during bond formation. This effect is evident in cross-coupling catalysis, where P(t-Bu)3-ligated species exhibit enhanced turnover for C-C bond formation due to the promoted elimination step. The identity of the participating ligands further modulates reductive elimination rates, with alkyl-aryl pairs generally undergoing faster elimination than hydride-alkyl pairs owing to differences in bond strengths and trans influences. In cis-oriented pairs, the trans ligand exerts a directing effect that facilitates coupling; for example, an trans to an alkyl promotes more rapid C-C elimination than a hydride in analogous positions, as observed in kinetic studies of Pd() dialkyl and alkyl complexes. This selectivity arises from the higher thermodynamic driving force and lower barriers for carbon-carbon coupling relative to carbon-hydrogen formation in late transition metals. Ligand architecture, beyond simple bulk, plays a crucial role through variations in type and design. Phosphines typically offer tunable steric profiles, while N-heterocyclic carbenes (NHCs) provide stronger σ-donation coupled with higher inherent steric demand, often quantified by the percent buried volume (%VBur), which correlates inversely with elimination barriers in some cases. For example, the bulky IPr NHC (1,3-bis(2,6-diisopropylphenyl)imidazolin-2-ylidene), with a %VBur of approximately 44%, slows C-H reductive elimination in Pd complexes by impeding the compact required for transfer, as demonstrated in computational and experimental analyses of NHC-ligated systems. In contrast, bidentate phosphines with wide bite angles (e.g., 102–110° in ligands like DPEphos or ) enforce cis orientation of the eliminating groups, stabilizing the near-tetrahedral and accelerating reductive elimination in catalytic cycles such as the . Steric factors also dictate selectivity in systems capable of competing eliminations. In mixed alkyl-aryl-hydride Pd(II) complexes, bulky ligands favor C-C over C-H reductive elimination by preferentially destabilizing pathways involving the more space-demanding hydride migration, thereby enhancing product yields in selective cross-couplings. Quantitative steric maps, derived from %VBur calculations, show that ligands exceeding 40% buried volume shift selectivity toward C-C formation, as the increased congestion hinders alternative H-transfer modes. These inherent steric and ligand influences can be combined with electronic tuning to optimize overall catalytic performance.

Structural and Geometric Effects

Reductive elimination rates are significantly influenced by the of the metal center, with 16-electron complexes generally proceeding faster than 18-electron species due to reduced steric congestion that allows closer approach of the ligands. For instance, in d⁸ square planar and systems, the unsaturated 16-electron configuration facilitates direct elimination, whereas octahedral 18-electron d⁶ complexes, such as alkyls, often require prior to access a five-coordinate before coupling. This preference manifests in pathways predominant in crowded 18-electron environments, where initial loss lowers the to promote the reaction, contrasted with rarer associative mechanisms that temporarily increase coordination. The geometric arrangement of ligands is essential, as reductive elimination mandates a orientation of the departing groups to enable optimal orbital overlap for new bond formation. In square planar d⁸ complexes, the inherent 90° angles support efficient elimination, while octahedral geometries necessitate distortions—such as toward square pyramidal—to align ligands properly and enhance reactivity. Theoretical analyses confirm that isomers are inactive, requiring prior , and highlight how geometric strain in distorted structures lowers activation barriers by improving heteroatom-carbon interactions in the . Structural features like chelating ligands profoundly modulate these processes by enforcing geometry and influencing fluxional behavior. Bidentate phosphines in octahedral (IV) dialkyl complexes, for example, form stable chelates that must undergo ring opening to generate a reactive five-coordinate for C-C bond formation, a step that is rate-limiting in such systems. In d⁸ square planar complexes, fluxionality enables rapid cis-trans interconversion via trigonal bipyramidal intermediates, ensuring access to the productive even from less favorable starting arrangements. X-ray crystallographic studies illustrate pre-elimination distortions that foreshadow reactivity, such as in bis(alkyl) complexes where steric bulk induces a widened C-Ni-C angle of 38.9° and elongated N-Ni bonds (2.055 and 1.982 ), deviating from ideal square planarity and facilitating elimination through strain relief. Computational investigations of transition states reveal optimal geometries, including P-Pt-P bite angles that expand by 4–6° from ground-state values (typically near 90°) during C-C coupling in diaryl systems, aligning the ligands for maximal overlap while elongating non-participating bonds. Regarding reversibility, higher coordination numbers in reactant complexes contribute to diminished stabilization of the lower-coordinate products, favoring irreversible elimination under typical conditions. This contrasts with three-coordinate intermediates in (III) or (II) systems, where reductive elimination can be reversible due to closer energetic proximity between starting materials and products.

External Activation Methods

External activation methods, such as photolysis and chemical , promote reductive elimination in organometallic complexes where is otherwise sluggish due to high activation barriers. These techniques introduce external energy or perturbations to facilitate ligand labilization or elevate the metal's , thereby enabling bond formation between cis-coordinated and reducing the metal center. Photolysis employs UV or visible light to induce reductive elimination by promoting ligand dissociation, often generating reactive 16-electron intermediates. In iridium complexes, irradiation leads to the reductive elimination of O₂, H₂, or HCl through initial labilization of bound ligands, forming low-coordinate species that accelerate the coupling step. For nickel(II) dialkyl complexes, visible light (wavelengths ≤427 nm) triggers C(sp³)–C(sp³) bond formation via Ni–C homolysis and radical rebound, with optimal reactivity at 370–390 nm yielding up to 79% product in sterically hindered systems. Wavelength dependence is pronounced, as shorter wavelengths (e.g., around 350–390 nm) enhance quantum yields by exciting metal-centered or ligand-to-metal charge transfer (LMCT) transitions, while longer wavelengths (>467 nm) show negligible activity. Chemical oxidation uses oxidants to temporarily increase the metal's , lowering the barrier for reductive elimination in reluctant complexes. Silver(I) salts, such as Ag⁺, serve as effective oxidants in by generating higher-valent Ni(III) or Ni(IV) species that undergo rapid . For instance, in Ni-mediated cross-couplings, oxidative promotion with Ag⁺ drives C–C bond formation from stalled Ni(II) intermediates, achieving yields up to 90% by facilitating two-electron reduction of the metal. Combined photo-oxidative approaches leverage synergy between light and oxidants, often via LMCT or metal-to-ligand charge transfer (MLCT) transitions, to enhance reductive elimination. In early systems like or diaryl complexes, LMCT excitation upon promotes followed by reductive elimination, yielding biaryl products through from ligands to metal d-orbitals. These methods can revive stalled catalytic cycles, but limitations include side reactions such as metal decomposition or unwanted pathways, particularly under prolonged or strong oxidants.

Synthetic Applications

Cross-Coupling Reactions

In cross-coupling reactions, reductive elimination constitutes the pivotal product-forming and catalyst-regenerating step, occurring after of an and with a to yield a metal complex bearing the two coupling partners. For instance, in the Suzuki-Miyaura reaction, this step transforms a (II) species coordinated to an from the halide and an alkyl or from the derivative into the zero-valent catalyst and the new C-C bond product, such as Ar-R. Prominent examples of cross-coupling reactions relying on reductive elimination include the Suzuki-Miyaura coupling for forming carbon-carbon bonds between aryl or halides and boronic acids, the Buchwald-Hartwig for C-N bond formation between aryl halides and amines, and the for C-C bonds using organozinc reagents with halides. In palladium-catalyzed aryl-aryl Suzuki-Miyaura couplings, reaction rates are notably enhanced by bases such as (K₃PO₄), which facilitates and thereby sets the stage for efficient reductive elimination. These reactions find broad industrial applications, particularly in pharmaceutical where precise C-C and C-N formations enable the of complex drug scaffolds, as seen in the production of kinase inhibitors and other therapeutics. Retention of during reductive elimination is achievable through chiral , preserving enantiopurity in asymmetric cross-couplings of prochiral substrates. A key challenge arises in alkyl-substituted variants, where competing β-hydride elimination from the alkyl-metal intermediate can divert the pathway toward or side products instead of the desired coupling. Ligand selection can accelerate reductive elimination to mitigate such issues.

Other Catalytic Transformations

Reductive elimination plays a pivotal role in C-H activation catalysis, where it typically occurs after the insertion of an unsaturated substrate into the metal-carbon bond generated by initial C-H oxidative addition, thereby releasing the functionalized product and regenerating the lower-valent metal catalyst. For instance, in the iridium-catalyzed borylation of aromatic C-H bonds, the cycle involves reaction of the iridium catalyst with a diboron reagent to form an iridium boryl species, followed by C-H activation and boryl transfer via reductive elimination to afford the arylboronate ester. This process, pioneered by Hartwig and coworkers, operates under mild conditions and achieves turnover numbers exceeding 10,000 for certain substrates, demonstrating the efficiency of reductive elimination in facilitating high catalytic activity. Rhodium-based variants of C-H borylation similarly rely on this step for C-B bond formation, often with turnover numbers above 10³ in transfer borylation protocols. In asymmetric C-H activation, reductive elimination can serve as the enantiodetermining event, as seen in rhodium(I)-catalyzed enantioselective functionalizations where the stereochemistry of the forming C-C or C-heteroatom bond is controlled by the chiral ligand environment during elimination. Recent advances include oxidatively induced reductive elimination (OIRE) for selective C-H bond functionalizations under mild conditions. In catalysis, reductive elimination contributes to chain propagation and transfer steps, particularly in transition-metal-mediated chain-growth mechanisms. Although in traditional olefin like Ziegler-Natta systems often proceeds via β-hydride elimination variants to generate alkenes and metal hydrides, reductive elimination variants appear in specialized systems, such as those involving migratory insertion followed by coupling to terminate or branch chains. Beyond these, reductive elimination features prominently in other transformations like hydrosilylation and decarbonylation. In standard hydrosilylation of alkenes, the step involves reductive elimination of the C-Si bond from an alkyl-metal-silyl intermediate, as exemplified in - and platinum-catalyzed additions of silanes to olefins, where it completes the after of the Si-H bond and olefin insertion. Certain variants, such as those in boryl-assisted processes, incorporate reductive elimination to form Si-H bonds, where a source facilitates elimination from a silyl-metal- , enabling dehydrogenative or transfer hydrosilylation pathways. In decarbonylation reactions, reductive elimination follows CO extrusion from acyl-metal intermediates; for instance, -catalyzed decarbonylation of acylsilanes proceeds via , CO release to generate an aryl , and subsequent reductive elimination with a coupling partner to yield biaryls or related products. Emerging applications draw inspiration from biological systems, where reductive elimination activates metal centers for challenging reductions. In synthetic mimics of Fe-nitrogenase, reductive elimination of H₂ from diiron hydride clusters clears coordination sites, enabling N₂ binding and stepwise reduction to , with computational and experimental studies confirming this step's role in overcoming kinetic barriers inherent to low-valent iron platforms. These bio-inspired processes highlight reductive elimination's potential in sustainable catalysis for . Overall, in these diverse transformations, reductive elimination parallels its role in cross-coupling by providing a facile route for product release and catalyst turnover.

References

  1. [1]
  2. [2]
    Uranium-mediated oxidative addition and reductive elimination
    Jun 23, 2015 · Oxidative addition, and its reverse reaction reductive elimination, constitute two key reactions that underpin organometallic chemistry and catalysis.
  3. [3]
    Reductive Elimination - an overview | ScienceDirect Topics
    Reductive elimination is defined as the second step in a reaction sequence where an initial adduct rearranges to form a new carbon–carbon bond, resulting in ...
  4. [4]
    Transition-metal-catalyzed reactions involving reductive elimination ...
    Jan 30, 2020 · Reductive elimination is a crucial bond-forming elementary reaction in various transition-metal mediated reactions.
  5. [5]
    [PDF] Reductive Elimination
    Reductive elimination, the reverse of oxidative addition, is most often seen in higher oxidation states because the formal oxidation state of the metal is ...Missing: definition seminal papers<|control11|><|separator|>
  6. [6]
    The Organometallic Chemistry of the Transition Metals
    Oxidative Addition and Reductive Elimination (Pages: 159-182) · Summary · PDF · Request permissions. CHAPTER 7. no. Insertion and Elimination ( ...
  7. [7]
    [PDF] PALLADIUM-CATALYZED CROSS COUPLINGS IN ORGANIC ...
    Oct 6, 2010 · Such a reductive elimination was demonstrated by A. Yamamoto in 1970 with the use of a diethylnickel complex.24 The nickel complex employed is ...
  8. [8]
    Whitesides Group Publications: The 1970s
    97. "Oxidative Addition and Reductive Elimination Reactions Involving ... 54. "Reactions of Alkylmercuric Halides with Sodium Borohydride in the ...Missing: migrations | Show results with:migrations
  9. [9]
    Group VIII Metal Complexes as Catalysts for ... - RSC Publishing
    Published on 01 January 1975. Downloaded by ... of oxidative addition and reductive elimination reactions ... Soc., 1966, 88, 3511; J. P. Collman, Accounts Chem.
  10. [10]
    Evidence for single metal two electron oxidative addition ... - Nature
    Dec 1, 2017 · Reversible single-metal two-electron oxidative addition and reductive elimination are common fundamental reactions for transition metals ...
  11. [11]
    C−H Activation for the Construction of C−B Bonds - ACS Publications
    The primary focus of this article is a review of issues of structure and bonding in boryl systems. Thus, the evidence for boryl ligands acting - to a greater ...Introduction · Rhodium-Catalyzed Borylation... · Iridium-Catalyzed Borylation of...
  12. [12]
  13. [13]
    Differences between the elimination of early and late transition metals
    Early transition metals (TMs), such as titanium, generally resist undergoing reductive elimination to form C–X bonds due to their weak electronegativity.
  14. [14]
    Structure-Reactivity Relationship in Allyl and 2-Propynyl Complexes ...
    during these reductive elimination reactions, an analysis of the detailed kinetic results in the case of the palladium com- plexes led us to suggest the ...
  15. [15]
    Electronic Dependence of C−O Reductive Elimination from ...
    These and related data point to the buildup of negative charge in the palladium-bound aryl group in the transition state for C−O reductive elimination.Abstract · Conclusions · Experimental Section
  16. [16]
    and Two‐Electron Unimolecular Reactions of Late 3d‐Metal ... - NIH
    Entropically, the radical dissociations are more favorable than the concerted reductive eliminations with their highly ordered transition structures.
  17. [17]
    [PDF] Reductive Elimination in Transition Metal Complexes - eDiss
    Sep 7, 2025 · Among the transition metals the reductive elimination generally proceeds faster for 3d metals than for the 4d metals, and faster for the 4d ...
  18. [18]
    Computational Insights into the Effect of Ligand Redox Properties on ...
    Mar 12, 2025 · We computationally investigated the Ar–Ar reductive elimination process in a series of biaryl Au(III), Pd(II), and Pt(II) complexes to explore the factors that ...<|control11|><|separator|>
  19. [19]
    Computed Ligand Electronic Parameters from Quantum Chemistry ...
    The best known is Tolman's electronic parameter (TEP) for phosphines, R3P, based on the position of the A1 ν(CO) vibration of (R3P)Ni(CO)3 in the IR spectrum.Missing: reductive | Show results with:reductive
  20. [20]
    In-depth insight into the electronic and steric effects of phosphine ...
    studied the reductive elimination of R–R (R = Me, Ph, t-Bu-C triple bond C–, t-Bu-C double bond C–) from a series of phosphine complexes of the type L2PR2 ...
  21. [21]
    Biaryl Phosphine Based Pd(II) Amido Complexes: The Effect of ...
    ... Pd(Ph)(NHPh). Thermolysis at 90 °C resulted in the loss of a PMe3 and ... For the reductive elimination of ethane from Pt(Me3I(PMePh2)2), the reported ...
  22. [22]
    [PDF] The Suzuki Reaction - Andrew G Myers Research Group
    Bulky phosphine ligands lead to a monoligated palladium species which is highly reactive to oxidative addition. These ligands are commercially available.
  23. [23]
    V Bur index and steric maps: from predictive catalysis to machine ...
    Dec 19, 2023 · In general, those studies found that ligands with larger %VBur values correspond to weaker binding to the metal centre, while ligands with ...
  24. [24]
    Reductive Elimination of Ethane from Five-Coordinate Platinum(IV ...
    The higher proportion of CH4 from 3b versus 3a is consistent with faster activation of an aryl versus an alkyl C−H bond, albeit steric constraints in the ...
  25. [25]
    Theoretical studies on C–heteroatom bond formation via reductive ...
    For the four-coordinate pathway the ease of reductive elimination to give M(PH3)2 and CH3–X follows the trend M = Pd < Pt < Ni. The reaction of the cis-M(PH3)2 ...
  26. [26]
    Reductive Elimination of d8-Organotransition Metal Complexes
    Jun 1, 1981 · A theoretical analysis of two aspects of the mechanism of reductive elimination is presented—how the choice of central metal and peripheral ...Missing: d10 | Show results with:d10
  27. [27]
    Reductive Elimination from Sterically Encumbered Ni–Polypyridine ...
    Sep 21, 2022 · Natural bond orbital analysis further confirmed a weaker overlap between the lone pairs of the nitrogen atoms and the p orbital of Ni in 4.
  28. [28]
    Force-modulated reductive elimination from platinum(II) diaryl ...
    Jul 26, 2021 · Here we quantify the rate of C(sp 2 )–C(sp 2 ) reductive elimination from platinum(II) diaryl complexes containing macrocyclic bis(phosphine) ligands.
  29. [29]
    Halide-Dependent Mechanisms of Reductive Elimination from Gold(III)
    Jun 11, 2015 · The PPh3-supported complexes can undergo both Caryl–X and Caryl–CF3 reductive elimination. Mechanistic studies of thermolysis at 122 °C reveal a ...Missing: octahedral | Show results with:octahedral
  30. [30]
    Photochemical reductive elimination of oxygen, hydrogen, and ...
    Photochemical reductive elimination of oxygen, hydrogen, and hydrogen chloride from iridium complexes ... The Photolysis of Chlorocarbonylbis ...
  31. [31]
    Unlocking Catalysis Using Oxidatively Induced Reductive Elimination
    Feb 17, 2025 · Among all the elementary reactions of transition metal organometallic chemistry, reductive elimination reactions allow the formation of a new σ- ...
  32. [32]
    A new era of LMCT: leveraging ligand-to-metal charge transfer ...
    Apr 17, 2024 · This work provides general criteria for designing transition metal complexes that exhibit low energy LMCT excited states and routes to drive photochemistry ...
  33. [33]
    Highly Active Catalyst for Suzuki–Miyaura Coupling to Form ...
    Jun 16, 2025 · To accelerate reductive elimination in Suzuki–Miyaura coupling (SMC) through secondary interactions with a Buchwald-type ligand, a ...
  34. [34]
    Mechanistic Aspects of the Palladium‐Catalyzed Suzuki‐Miyaura ...
    Jul 16, 2021 · The reaction is only slightly affected by steric hindrance and has a wide tolerance for a broad range of functional groups. ... More importantly, ...
  35. [35]
    Cross-Coupling and Related Reactions: Connecting Past Success ...
    ... reductive elimination, are similar to those proposed in traditional cross-coupling reactions. The key vinyl boronate compounds in conjunctive cross-coupling ...
  36. [36]
    Insights into the elementary steps in Negishi coupling through ...
    Jun 18, 2012 · Reductive elimination is an essential step for bond formation. Similar to oxidative addition, most of the knowledge about reductive elimination ...
  37. [37]
    Transition metal–catalyzed alkyl-alkyl bond formation - Science
    Apr 14, 2017 · An impediment to alkyl-alkyl cross-coupling: β-hydride elimination. As mentioned above, most early studies of cross-couplings used palladium ...Structured Abstract · Primary Alkyl Electrophiles · Secondary Alkyl...