Fact-checked by Grok 2 weeks ago

Oxidative addition

Oxidative addition is a fundamental reaction in in which a low-valent complex inserts into a bond of a , such as H₂, an alkyl (R–X), or a C–H bond, forming two new metal– bonds while simultaneously increasing the metal's formal , , and d-electron count by two units. The concept was established in the 1960s through work by researchers such as Vaska, Wilkinson, and others. This process typically requires a coordinatively unsaturated metal center, often a 16-electron or one generated by ligand dissociation from an 18-electron complex, and is most common with d⁸ or d¹⁰ metals like Pd(0), Rh(I), Ir(I), or Pt(0). The reaction is reversible under appropriate conditions and serves as the microscopic reverse of . Mechanistically, oxidative addition proceeds through diverse pathways depending on the substrate and metal. Concerted mechanisms involve a three-center for non-polar substrates like H₂, as seen in the addition to (IrCl(CO)(PPh₃)₂) to yield a cis-dihydride Ir(III) . For polar substrates like alkyl halides, Sₙ2 pathways predominate, featuring nucleophilic attack by the metal at carbon with inversion of configuration, while radical mechanisms occur via single-electron transfer, often in non-chain or chain processes. Ionic mechanisms, such as stepwise followed by anion , apply to substrates like HX. These variations highlight the reaction's versatility, with rates influenced by factors like solvent polarity, metal electronics, and sterics—electron-rich metals and polar solvents accelerate the process. Oxidative addition plays a pivotal role in by activating unreactive bonds, enabling transformations central to synthetic chemistry. It is the initial step in catalytic cycles for using (RhCl(PPh₃)₃), , and cross-coupling reactions like the or Heck processes, where Pd(0) adds to aryl halides. In C–H activation, it facilitates the conversion of alkanes or arenes into functionalized products, as exemplified by or complexes inserting into C–H bonds. Its study has driven advances in , underscoring its status as an for organometallic reactivity.

Introduction

Definition and characteristics

Oxidative addition is a key in in which a neutral substrate, denoted as XY, adds to a low-valent center M, forming the product M(X)(Y). This process simultaneously increases the of the metal by two units and its by two, while the d-electron count of the metal decreases by two electrons. The general reaction can be represented as: \mathrm{M^{n} + XY \rightarrow M^{n+2}(X)(Y)} A prerequisite for oxidative addition is the availability of a vacant coordination site on the metal complex, which may arise from prior dissociation. The reaction is particularly favored for electron-rich, low-valent complexes of late transition metals, such as d^{10} species like Ni(0), Pd(0), and Pt(0), or d^8 species like Rh(I) and Ir(I), where the higher oxidation states are accessible and stable. Common examples include the conversion of Ni(0) to Ni(II) upon addition of alkyl halides, Pd(0) to Pd(II) with aryl halides, and Ir(I) to Ir(III) with dihydrogen or alkyl halides. The reversibility of oxidative addition, whose microscopic reverse is , underscores its thermodynamic accessibility in many systems, particularly when the added ligands are weakly bound. Transition metals' ability to adopt multiple oxidation states facilitates this two-electron process at the metal center. Furthermore, oxidative addition often transforms 16-electron complexes into 18-electron species, adhering to the , which posits that organometallic complexes with 18 valence electrons around the metal achieve enhanced stability analogous to the in main-group chemistry.

Historical development

The concept of oxidative addition emerged in the early through empirical observations of complexes reacting with small molecules to form higher-oxidation-state products. Lauri Vaska and John W. DiLuzio first reported the square-planar iridium(I) complex IrCl(CO)(PPh₃)₂, known as , in 1961, which underwent oxidative addition with acids to yield stable hydrido-carbonyl species. In 1962, Vaska reported the oxidative addition of hydrogen to the same complex, marking one of the earliest documented examples. In 1963, Vaska demonstrated the reversible oxidative addition of dioxygen to the same complex, forming a dioxygen adduct that highlighted the potential for binding and activation of inert molecules. These findings established oxidative addition as a key reactivity pattern for low-valent, coordinatively unsaturated metal centers, particularly d⁸ square-planar species. In the mid-1960s to , researchers expanded the scope and mechanistic understanding, linking oxidative addition to . Joseph Chatt and Brian L. Shaw reported in 1965 the oxidative addition of methyl iodide to (0) complexes, providing evidence for alkyl activation in catalytic cycles. James P. Collman popularized the term "oxidative addition" in 1968 while studying and systems, emphasizing its role in generating reactive intermediates for and processes. Jack Halpern's kinetic studies in the further elucidated two-electron mechanisms for alkyl additions to and complexes, demonstrating rate dependencies on substrate and metal nucleophilicity. Concurrently, Chadwick A. Tolman integrated oxidative addition into the framework in 1972, explaining how 16-electron precursors favor addition to achieve stable 18-electron products, shifting the field from isolated reactions to predictive models. By the 1980s, mechanistic investigations refined the pathways, with George M. Whitesides' work on and establishing SN2-like mechanisms for polar substrate additions to and centers. Studies by Charles E. Johnson and Eisenberg in 1985 revealed radical chain aspects in rhodium-catalyzed hydrogenations, where oxidative addition of H₂ initiated . These developments underscored the versatility of oxidative addition, from concerted to stepwise processes, as reviewed comprehensively by Jay A. Labinger in 2015, who traced its evolution into a cornerstone of organometallic reactivity.

Role in organometallic chemistry

Significance in catalysis

Oxidative addition plays a central role in catalytic cycles, frequently serving as the initial step in cross-coupling reactions and C-H activation processes, where it coordinates with to construct new carbon-carbon or carbon-heteroatom bonds. This transformation enables low-valent metal centers, such as Pd(0) or Rh(I), to insert into otherwise unreactive bonds like C-H or C-X, thereby activating substrates for subsequent reactivity that cannot be achieved through conventional organic synthetic routes. In prototypical cross-coupling cycles, oxidative addition initiates the activation of electrophilic partners. For instance, in the Suzuki-Miyaura reaction, a Pd(0) precatalyst undergoes oxidative addition with an to generate a Pd(II) aryl intermediate, setting the stage for with an organoborane and ultimate to form the coupled biaryl product. Similarly, the begins with oxidative addition of an to Pd(0), followed by coordination and insertion of an , leading to the formation of a substituted upon β-hydride elimination and regeneration of the catalyst. These cycles exemplify how oxidative addition drives efficient bond-forming processes under mild conditions. The significance of oxidative addition extends to broader synthetic advancements, facilitating asymmetric catalysis through the integration of chiral ligands that impart stereocontrol during the addition step, thus enabling enantioselective construction of complex molecules. Additionally, it supports objectives by promoting the use of earth-abundant feedstocks, such as simple alkyl or aryl halides, and minimizing waste through atom-economical cycles that reduce reliance on stoichiometric oxidants or protecting groups. Reviews highlight its involvement as a key elementary step in the majority of homogeneous catalytic processes, underscoring its foundational impact on modern synthetic methodology.

Comparison with reductive elimination

Reductive elimination is the microscopic reverse of oxidative addition, in which a higher-oxidation-state organometallic complex of the form M^{n+2}(X)(Y) expels the ligands X and Y as the neutral molecule XY, thereby decreasing the metal's oxidation state to n and its coordination number by two. Key differences between the two processes arise from their electronic and steric requirements. Oxidative addition is typically favored by electron-rich, low-oxidation-state metals that can accommodate additional electron density from the incoming ligands, as well as by substrates featuring polar or relatively weak bonds. In contrast, reductive elimination is promoted by electron-poor metals in higher oxidation states, where the metal-ligand bonds are weakened, and by sterically crowded cis-oriented ligands that drive the intramolecular coupling of X and Y. Thermodynamically, oxidative addition and are often reversible, with the position of equilibrium governed by factors such as the metal's identity, supporting ligands, and solvent environment; for example, the change (ΔG) is modulated by the at the metal center and the relative bond strengths of the M–X, M–Y versus X–Y bonds. In palladium-catalyzed cross-coupling reactions, such as the Suzuki–Miyaura coupling, oxidative addition of an (Ar–X) to a Pd(0) species generates a Pd(II) intermediate, which, after with an organoborane, undergoes to afford the coupled Ar–R product and regenerate Pd(0). A distinctive feature of reductive elimination is the migratory aptitude of the ligands involved, which dictates selectivity when multiple coupling pathways are possible; for instance, alkyl ligands often display higher migratory aptitude than aryl ligands in certain systems, thereby influencing the directionality and product distribution within catalytic cycles.

Reaction scope and factors

Suitable substrates

Oxidative addition reactions typically involve the cleavage of homonuclear or heteronuclear bonds in substrates that coordinate to low-valent centers, increasing the metal's and by two units. Common bond types include H-H (dihydrogen), H-X (where X is a ), C-H, and C-X (in alkyl or aryl halides). These reactions are facilitated by the formation of strong metal-ligand bonds that compensate for the energy required to break the bond. Diatomic molecules such as H₂ and Cl₂ serve as classic substrates, with H₂ undergoing oxidative addition to complexes like Vaska's IrCl(CO)(PPh₃)₂ to form cis-dihydrido species. Polar substrates like HCl and CH₃I are also highly reactive; for instance, HCl adds ionically to or centers in polar solvents, while CH₃I reacts with methylplatinum complexes via an Sₙ2 pathway. Nonpolar substrates, such as alkanes, participate through C-H activation, where sp³ C-H bonds in or add to metals like or , often requiring high temperatures due to the bond's relative strength. Bond dissociation energies (BDEs) play a key role in substrate suitability, with lower BDE values generally favoring reactivity. The H-H bond in dihydrogen has a BDE of 436 kJ/mol, making it one of the easiest to cleave, while H-Cl stands at approximately 431 kJ/mol and C-H in alkanes around 410 kJ/mol. In contrast, C-X bonds vary significantly: C-I in CH₃I is weaker at 234 kJ/mol, enabling facile addition, whereas C-F in CH₃F has a high BDE of 452 kJ/mol, which limits oxidative fluorination reactions. Aromatic C-H bonds, with BDEs around 472 kJ/mol for , are more reactive than aliphatic counterparts due to the stability of the resulting aryl-metal hydride. Less common substrates include O₂, SO₂, N-N, and Si-H bonds. Molecular oxygen adds to square-planar Ir(I) complexes like Vaska's compound, forming peroxo-Ir(III) species, while SO₂ undergoes reversible η²-S,O coordination and oxidative addition to four-coordinate metal centers such as Pd(0) or Pt(0). N-N bonds, particularly in azo compounds with N=N double bonds, can participate in multi-electron oxidative additions to early transition metals like , though this is rare. Si-H bonds in silanes, with BDEs around 384 kJ/mol for PhSiH₃, add readily to Pd(0) or Pt(0) centers, forming silyl-hydride complexes essential for hydrosilylation catalysis. Oxidative addition to C-C bonds is uncommon due to their high BDEs (typically >350 kJ/mol) and unfavorable , and the process invariably proceeds with stereochemistry, preserving the relative geometry of the added fragments.

Influencing factors

The reactivity of oxidative addition is profoundly influenced by the electronic and steric properties of the metal center. Low-valent, electron-rich d^{10} metals such as Pd(0) and Pt(0) are particularly amenable to this process due to their ability to readily increase in by two units while accommodating additional ligands, with reactivity correlating positively with more negative oxidation potentials that facilitate to the . For instance, Pd(0) complexes exhibit enhanced rates when the metal's is high, as measured by molecular electrostatic potential at the Pd , lowering activation barriers for addition. Ligand effects play a critical role in modulating both the rate and pathway of oxidative addition. π-Acceptor ligands, such as , can accelerate the reaction by labilizing other ligands and creating coordination vacancies, particularly in dissociative mechanisms, though they may compete for binding sites and slow rates in associative paths. Steric bulk from ligands like P(t-Bu)_3 promotes low-coordination (e.g., monoligated [Pd(0)L]), favoring dissociative oxidative addition to sterically hindered substrates by reducing the energy barrier for ligand dissociation. In contrast, less bulky or electron-donating ligands like PCy_3 stabilize bis-ligated complexes, directing the reaction through associative bimolecular pathways. Solvent polarity and reaction conditions further control the feasibility and selectivity of oxidative addition. Polar aprotic solvents, such as DMF or , stabilize charged transition states and favor ionic or S_N2-type , increasing rates for polar like alkyl halides. Elevated temperatures are often necessary for endothermic additions, such as those involving unactivated C-H bonds, to overcome kinetic barriers, with rates showing Arrhenius dependence on input. Stereochemical outcomes depend on the : concerted additions typically proceed with retention of at the substrate, while S_N2 pathways result in inversion. Quantitative analyses reveal systematic trends in reactivity. For aryl substrates, Hammett parameters show positive ρ values, indicating that electron-withdrawing substituents accelerate oxidative addition by stabilizing the developing positive charge on the aryl ring during C-X bond cleavage. Rate laws for bimolecular oxidative additions are typically second-order overall, in both and concentrations, as observed in phosphine-supported Pd(0) systems with aryl bromides.

Mechanisms of oxidative addition

Concerted mechanism

The concerted of oxidative addition proceeds via a synchronous, two-electron process involving a three-center , in which the metal center inserts directly into the X-Y bond of the , simultaneously breaking the X-Y bond and forming two new metal-ligand bonds. This pathway is characteristic of nonpolar substrates, such as dihydrogen (H₂) and certain alkyl halides lacking β-hydrogen atoms, where polar or stepwise mechanisms are disfavored due to the symmetric and nonpolar nature of the bond. In this , the reaction maintains cis in the product, reflecting the concerted insertion without inversion or steps. A seminal example is the oxidative addition of H₂ to Vaska's complex, trans-IrCl(CO)(PPh₃)₂, which produces the cis-dihydride complex cis-IrH₂Cl(CO)(PPh₃)₂. This reaction exemplifies the pathway for dihydrogen activation: \text{trans-IrCl(CO)(PPh}_3)_2 + \text{H}_2 \rightarrow \text{cis-IrH}_2\text{Cl(CO)(PPh}_3)_2 The square-planar iridium(I) starting complex adopts an octahedral geometry in the product, with the two hydride ligands adding cis to the metal center. Experimental evidence supporting the concerted nature includes second-order kinetics (first-order in both metal complex and H₂ concentration) and a small kinetic isotope effect (KIE = k_H/k_D ≈ 1.1 at 25°C), indicating that H-H bond cleavage does not occur in the rate-determining step but rather in a symmetric transition state. Linear free energy relationships, such as those correlating reaction rates with electronic variations in phosphine ligands, further corroborate the synchronous insertion without significant charge separation. Density functional theory (DFT) computations provide additional insight, revealing a weakly bound σ-complex intermediate (η²-H₂) preceding the , where the oxidative addition proper occurs through a tight three-center structure. In this , the H-H bond elongates slightly, facilitating the insertion. Recent analyses emphasize the role of orbital interactions in stabilizing the : the metal d_{z^2} orbital donates electrons to the σ bonding orbital of H₂, while back-donation from the filled metal d orbital to the σ* antibonding orbital of H₂ weakens the H-H bond, enabling concerted cleavage. These interactions, highlighted in 2021 reviews, underscore the pericyclic-like character of the process for nonpolar bonds, distinguishing it from stepwise pathways.

SN2-type mechanism

The SN2-type in oxidative addition proceeds via a nucleophilic at the carbon center of a polar C-X , where the low-valent metal acts as the . In this pathway, the metal attacks the carbon from the backside, leading to inversion of and of the C-X , with concomitant formation of M-C and M-X σ-bonds. This two-electron process involves a with partial positive charge on carbon and negative charge on the departing , distinguishing it from synchronous concerted additions to nonpolar bonds. This mechanism is well-suited to polar substrates such as primary alkyl (e.g., CH₃I) and benzyl (e.g., ), where the electrophilic carbon is accessible and the C-X bond is polarized. A key example is the oxidative addition of CH₃I to Pd(0) species, such as [Pd(PtBu₂Me)₂], yielding the Pd(II) complex [Pd(CH₃)(I)(PtBu₂Me)₂]; the reaction follows second-order with the rate law rate = k [Pd(0)][CH₃I], consistent with a direct bimolecular interaction. Supporting evidence includes stereochemical inversion at the chiral carbon in reactions with optically active alkyl , mirroring classic organic SN₂ displacements. The leaving group order I > Br > Cl reflects the ease of halide departure, with reacting fastest due to its weaker C-I bond and better . Protic solvents accelerate the process by solvating the developing ionic character in the , enhancing rates compared to aprotic media. Variations of this mechanism can involve an initial associative step, where the metal coordinates to the C-X forming an η²-alkyl before the nucleophilic displacement occurs, particularly with electron-rich Pd(0) or Pt(0) .

Ionic mechanism

The ionic of oxidative addition proceeds via the initial of the XY into free ions X⁻ and Y⁺, followed by their sequential coordination to the metal M, resulting in an overall two-electron oxidation of the metal. This stepwise process typically begins with the addition of the more electrophilic fragment (often Y⁺) to form an intermediate cationic , such as MH⁺ from H⁺ coordination, with the anionic fragment (X⁻) binding subsequently to yield the final neutral or ionic product. This pathway is particularly suited to highly polar substrates that readily dissociate, such as strong acids like HCl, which ionize to H⁺ and Cl⁻ in solution. The reaction is favored for low-valent, electron-rich late complexes, though it is more prevalent in early transition metals due to their higher oxophilicity toward cations. A representative example is the addition of HCl to a rhodium(I) complex, such as trans-[RhCl(CO)(PPh₃)₂], yielding a rhodium(III) species [RhCl₂H(CO)(PPh₃)₂], often observed in protic solvents where dissociation is enhanced. Kinetic studies support this mechanism through first-order dependence on substrate concentration, indicative of rate-determining initial addition, and negative activation entropies consistent with associative ion-pair formation in the . Ion-pairing effects further corroborate the pathway, as added anions accelerate the reaction by stabilizing the cationic and enhancing the metal's nucleophilicity. These observations are prominent in polar or protic solvents, which promote substrate and solvation, whereas nonpolar media suppress the process. Despite its utility for protic substrates, the ionic mechanism is relatively rare overall, owing to the high dissociation energies required for most XY bonds, limiting it primarily to hydrogen halides and a few highly electrophilic species in early metals. In contrast to concerted nucleophilic pathways like SN2-type mechanisms for covalent polar substrates, the ionic route involves fully dissociated species and discrete intermediates, emphasizing solvent-mediated ion separation.

Protonative mechanism

The protonative mechanism describes a heterolytic pathway for oxidative addition, primarily involving the of O-H or N-H bonds in protic substrates like alcohols and amines. In this process, a low-valent metal center, such as Pt(0) or Pd(0), first coordinates to the on the oxygen or nitrogen atom of the substrate. This coordination orients the X-H bond (X = O, N) proximal to the metal, enabling a subsequent proton transfer step mediated by an assisting species, typically or an molecule itself. The transfer proceeds via a cyclic , distinguishing this mechanism from sequential ionic pathways by its concerted proton movement within the cycle, ultimately yielding a trans metal-hydride-X product and increasing the metal's and by two. A representative example occurs with as the , where Pt(0) complexes, such as Pt[P(i-Pr)3]3, react in aqueous media to form trans-[Pt(H)(OH)(P(i-Pr)3)2]. Here, the pathway begins with water coordination to the 14-electron Pt(0) species, followed by base-assisted through a four-center cyclic involving Pt, the oxygen of water, the proton, and the base. This generates a highly hydroxo species capable of catalyzing like the water-gas shift or nitrile hydration. The geometry arises from the anti addition inherent to the cyclic , and the process is favored in protic environments due to the of the O-H bond. For N-H bonds, the protonative mechanism similarly activates , with 2010s studies highlighting its role in Pd(0)-catalyzed processes. For instance, adds across a Pd(0) center bearing a cyclic (alkyl)(amino) via a cyclic incorporating two molecules—one coordinating and the other facilitating proton —yielding trans-[Pd(H)(NHPh)(CAAC)(py)]. Computational modeling using (PBE0-D3BJ/def2-TZVPP) confirms this, showing a rate-determining cyclic with a barrier of approximately 108 kJ/mol for analogous O-H additions, and low overall barriers (ca. 20-30 kJ/mol) for N-H cases, supporting facile reactivity without external bases. These models also reveal stabilizing interactions akin to agostic bonding in the pre-coordination step, where partial C-H or lone-pair donation aids approach, though the core proton remains heterolytic. This mechanism's prevalence in amine activations underscores its utility in additive-free , often leading to trans products that influence downstream selectivity in or reactions.

Radical mechanism

In the radical mechanism of oxidative addition, a low-valent metal center engages in single-electron transfer (SET) with a substrate such as an alkyl or aryl halide (XY), generating a metal radical cation (M⁺•) and a substrate radical anion (XY⁻•). These odd-electron species then undergo rapid coupling within a solvent cage to yield the two-electron oxidative addition product M(X)(Y), often via a geminate radical pair. Alternatively, a radical rebound pathway may operate, wherein the substrate dissociates into X• and Y• fragments that recombine with the metal center. This pathway is particularly relevant for substrates like aryl halides (e.g., PhBr) under non-standard conditions involving radical initiators such as peroxides, (AIBN), or light, which promote homolytic cleavage or SET. Unlike polar mechanisms, the radical process favors electron-poor metals and is accelerated by basic ligands that facilitate initial electron donation. A key example is the reaction of Pt(0) complexes, such as Pt(PPh₃)₄, with (PhCH₂Br), where free s are implicated as intermediates, detected by electron spin resonance (ESR) . Supporting evidence includes inhibition by radical scavengers such as galvinoxyl or , which suppress product formation by quenching intermediates, and chemically induced (CIDNP) effects in NMR spectra indicative of radical pair recombination during oxidative addition to Pd(0) or Pt(0) centers. The radical mechanism remains controversial, often viewed as a side pathway competing with concerted or SN2-type processes rather than the dominant route under standard conditions; early studies highlighted its role in homolytic displacements, while recent DFT analyses clarify that it predominates only with specific initiators or in aprotic solvents. Variations appear in photocatalysis, where visible light excites a to generate radicals that facilitate SET to the metal, enabling oxidative addition in systems like Ni-catalyzed cross-couplings or Cu-mediated arylations, bypassing traditional two-electron pathways.

Applications

Synthetic transformations

Oxidative addition is a cornerstone of laboratory-scale , particularly in palladium-catalyzed cross-coupling reactions that forge carbon-carbon and carbon-nitrogen bonds from readily available precursors. This elementary step, involving the insertion of a low-valent metal center into a bond, generates reactive organometallic intermediates that enable selective transformations under mild conditions, often at temperatures below 100°C with high . These reactions have revolutionized the construction of complex molecules, including pharmaceuticals and natural products, by allowing the replacement of traditional multi-step sequences with convergent assemblies. In the Buchwald-Hartwig amination, oxidative addition of Pd(0) to an aryl bromide (ArBr) forms an aryl-Pd(II)-Br species, which undergoes coordination and subsequent to yield arylamines (Ar-NR₂). Developed independently by Buchwald and Hartwig in the mid-1990s, this method accommodates primary and secondary , including those with sensitive functional groups, and typically delivers 80-95% yields using bulky monophosphine ligands like or DavePhos in or dioxane at 80-100°C. For instance, the coupling of with proceeds smoothly to form in 92% yield under these conditions. The exemplifies halide insertion via oxidative addition, where Pd(0) adds across an organic halide (R'-X) to generate R'-Pd(II)-X, followed by with an organozinc reagent (R-ZnX') and to produce R-R'. Introduced by Negishi in 1977, this transformation excels in alkyl-aryl bond formation, tolerating a broad range of functional groups, and achieves 85-95% yields under mild conditions with phosphine-ligated Pd catalysts, such as in the synthesis of 1-phenylpropane from iodobenzene and ethylzinc chloride. Sonogashira and Stille couplings further highlight oxidative addition's versatility in C-C bond formation. In the Sonogashira reaction, Pd(0) oxidative addition to an aryl or vinyl halide precedes transmetalation with a copper(I) acetylide, yielding enynes (Ar-C≡C-R) in 90-98% yields typically at room temperature to 60°C with Pd(PPh₃)₄ and CuI in amine solvents; a representative example is the coupling of iodobenzene with phenylacetylene to form diphenylacetylene in 95% yield. The Stille coupling involves similar oxidative addition to the halide, followed by transmetalation with an organostannane (R-SnR'₃), enabling stereoretentive transfers and 80-95% yields under ligand-free or phosphine-supported conditions, as seen in the preparation of stilbene from tributylstannylbenzene and iodobenzene. Beyond traditional cross-couplings, oxidative addition facilitates C-H activation for arylation, where oxidative addition of Pd(0) to an arylating agent like an or generates an organopalladium(II) species, which then undergoes C-H activation (e.g., via palladation or concerted metallation-deprotonation with directing group assistance) to enable direct C-H to C-Ar conversion. Fagnou's methodology for intramolecular C-H arylation of arenes with aryl halides proceeds via this pathway, affording fused biaryls in 70-90% yields at 110°C with Pd(OAc)₂ and PCy₃, such as the cyclization of 2-bromobiphenyl to fluorene in 85% yield. Following oxidative addition, subsequent migratory insertion steps expand synthetic utility, as in olefin insertion where the alkyl- or aryl-Pd(II) intermediate coordinates and inserts an , leading to β-aryl carbonyls or styrenes after β-hydride elimination. This sequence underpins variants of the , achieving 80-95% yields with chiral ligands for enantiocontrol. Enantioselective oxidative additions are enabled by chiral ligands like in Pd systems, which impose asymmetry during the coupling cycle to produce enantioenriched products with ee >90%. For example, in asymmetric Negishi couplings of secondary alkylzinc reagents with aryl bromides, (R)- delivers the products in 88% yield and 95% ee under mild conditions, highlighting the precision achievable in stereoselective .

Industrial catalysis

Oxidative addition plays a pivotal role in several large-scale industrial catalytic processes, enabling the efficient transformation of basic feedstocks into commodity chemicals through transition metal catalysis. One of the most prominent examples is the for the synthesis of acetic acid from and , developed by (now part of ) in the 1970s. In this rhodium-iodide catalyzed reaction, the oxidative addition of methyl iodide (CH₃I) to the Rh(I) center is the rate-determining step, forming a Rh(III)-methyl intermediate that undergoes migratory insertion with CO to yield the acetyl species, ultimately producing acetic acid after iodide . This process operates under mild conditions (150–180°C, 30–40 bar), achieving selectivities over 99%, and has been a cornerstone of acetic acid production, with global capacity exceeding 20 million metric tons annually as of 2023, valued at billions of dollars in the . Hydroformylation, or oxo synthesis, represents another cornerstone application, where rhodium-phosphine catalysts facilitate the addition of (H₂/CO) to olefins, with oxidative addition of H₂ to the Rh(I) species being a critical step in forming the active dihydride intermediate for production. Developed by (now Dow) and Ruhrchemie in the 1970s, low-pressure rhodium systems have revolutionized the process, enabling high for linear aldehydes used in detergents and plastics; annual global production exceeds 10 million tons as of 2023, underscoring its economic impact as one of the largest homogeneous catalytic processes. In , oxidative addition underpins palladium-catalyzed Suzuki-Miyaura cross-couplings on ton-scale for synthesizing active pharmaceutical ingredients, such as losartan and , where the oxidative addition of aryl halides to Pd(0) initiates the C-C bond formation with boronic acids. These reactions have enabled efficient, scalable production of complex molecules, with industrial implementations achieving high yields and supporting billions in annual pharmaceutical revenue. Despite these successes, industrial applications face challenges in catalyst recovery and scalability, particularly for homogeneous systems where rhodium and palladium are expensive and prone to leaching. Techniques like ligand modification and biphasic solvents have been adopted to improve separation, as seen in hydroformylation plants, but ongoing efforts are needed to minimize losses and enhance sustainability without compromising activity.

Recent advances

Mechanistic insights post-2020

A comprehensive 2021 review synthesized the historical development and mechanistic diversity of in , emphasizing its role in and highlighting pathways such as concerted, SN2-type, and mechanisms while noting emerging observations in f-block elements. In 2025, researchers at Penn State University reported a "flipped" of H₂ to (II) and (II) complexes, where heterolysis of H₂ delivers electrons from the substrate to the metal center—contrary to the conventional metal-to-substrate electron donation—followed by protic rebound to form hydrido complexes, resulting in inverted compared to traditional pathways. Density functional theory (DFT) calculations in 2025 from the laboratory demonstrated that external electric fields at electrode interfaces can selectively stabilize versus in surface-bound Pd catalysts, enabling reversible control over reaction equilibria with field strengths as low as 0.1 V/Å. Concurrently, single-electron transfer (SET) mechanisms in were clarified through studies showing that SET initiates of unactivated alkyl halides via radical intermediates that mimic oxidative addition, bypassing two-electron barriers and achieving high selectivity under mild conditions. Experimental evidence from ultrafast in 2023 resolved dynamics in Rh-catalyzed C-H activation, revealing lifetimes on the order of 100 for the oxidative addition step and confirming concerted two-electron character through orbital-specific shifts. These insights have resolved longstanding debates between (SET) and two-electron mechanisms in cross-coupling reactions, with 2024 studies mapping branching ratios for addition to (0) species—up to 40% pathway under certain ligands—via combined electroanalytical and computational analysis, enabling predictive control over selectivity.

Novel catalytic systems

Recent advancements in catalytic systems exploiting oxidative addition have introduced innovative approaches that enhance efficiency and expand substrate scope beyond traditional frameworks. Single-atom catalysts, developed through an "anchoring-borrowing" strategy combined with facet engineering, enable cross-coupling reactions by defying conventional oxidative addition prerequisites, allowing activation of challenging C-Cl bonds under mild conditions. These artful single-atom catalysts (ASACs) demonstrate superior stability and selectivity, facilitating aryl couplings with turnover numbers exceeding 10,000. Complementing this, on-demand generation of oxidative addition complexes (OACs) from stable organopalladate salts has streamlined C-C coupling processes. By simply mixing the salt with mono- or bidentate ligands, these complexes form rapidly at , serving as precatalysts for Suzuki-Miyaura and Negishi couplings with diverse aryl and alkyl halides, achieving yields up to 98% without preactivation steps. This method reduces synthetic complexity and waste, promoting greener protocols for pharmaceutical intermediates. In C-H functionalization, novel (I)/(III) catalysts have emerged as key enablers, leveraging oxidative addition to alkyl halides for selective sp³ C-H . These mononuclear systems, characterized by advanced spectroscopic techniques, support directed functionalizations under ambient conditions, converting unactivated s to value-added products like alkylated heterocycles with site selectivity greater than 90%. Similarly, electric field-tuned systems allow reversible oxidative addition/ at interfaces, enabling dynamic control in molecular switches and adaptive , where applied fields modulate reaction equilibria to achieve up to 5-fold rate enhancements. Extensions to s-block metals represent a frontier in mimicking oxidative addition, with rare Mg(0)-mediated C-H activations reported in asymmetric complexes. These low-valent magnesium species undergo formal oxidative addition to aryl iodides, forming intermediates that facilitate hydrofunctionalizations with enantioselectivities reaching 99% , offering earth-abundant alternatives to precious metals under milder conditions. Calcium analogs, though less developed, show analogous reactivity in mixed aggregates, broadening s-block utility in selective couplings. Overall, these innovations yield higher selectivity and operational simplicity, exemplified by asymmetric C-C couplings attaining 99% at and .

References

  1. [1]
    [PDF] THE ORGANOMETALLIC CHEMISTRY OF THE TRANSITION ...
    This book covers organometallic chemistry of transition metals, including Werner complexes, the trans effect, metal alkyls, carbonyls, and the 18-electron rule.
  2. [2]
    [PDF] Oxidative Addition
    Oxidative additions are very diverse mechanistically, and we therefore consider each type separately. 1. Concerted, or three-center, oxidative addition ...Missing: review | Show results with:review
  3. [3]
    OA.2. Reaction Overview - oxidative addition
    Oxidative addition is a general term for the insertion of a metal between two atoms that were previously bonded together.Missing: definition | Show results with:definition
  4. [4]
  5. [5]
    14.1.2: Oxidative Addition - Chemistry LibreTexts
    May 3, 2023 · An oxidative addition is a reaction in which the coordination number and the oxidation state of the metal ion both increase by 2.
  6. [6]
    [PDF] Oxidative Addition
    Oxidative addition is usually favored by strongly donating ligands (electron rich substrates) because these stabilize the increased oxidation state of the ...
  7. [7]
    Oxidative Addition | OpenOChem Learn
    Oxidative Addition. Oxidative addition reactions involve increasing both the coordination number and the oxidation state of the metal by two.
  8. [8]
    4.7: Oxidative Addition- General Ideas - Chemistry LibreTexts
    Aug 28, 2022 · The metal must have a stable oxidation state two units higher than its current OS for oxidative addition to occur. For the reaction to work well ...Missing: prerequisites rule<|control11|><|separator|>
  9. [9]
    Journal of the American Chemical Society
    Article June 1, 1961. CARBONYL AND HYDRIDO-CARBONYL COMPLEXES OF ... oxidative addition of 1-iodopolyynes to Vaska's complex. Dalton Transactions ...Missing: original | Show results with:original
  10. [10]
    We Already Know Everything about Oxidative Addition to Pd(0)
    Aug 14, 2023 · The oxidative addition (OA) of organic electrophiles to Pd(0) is a fundamental step in organopalladium chemistry and plays a key role in ...
  11. [11]
    Oxidative Addition Reactions - SpringerLink
    Two-electron oxidative addition reactions are defined by equation (1). In this reaction where the A-B bond is activated, the metal center M is formally ...Missing: coined | Show results with:coined
  12. [12]
    The continuum of carbon–hydrogen (C–H) activation mechanisms ...
    Dec 10, 2021 · Oxidative addition, one of the most studied fundamental transformations in organometallic chemistry, is generally reserved for low valent ...
  13. [13]
    Emerging Trends in Cross-Coupling: Twelve-Electron-Based L1Pd ...
    Oct 3, 2022 · Computational, kinetic, and experimental studies indicate that all three steps in the catalytic cycle─oxidative addition, transmetalation, and ...
  14. [14]
    Asymmetric Catalytic Aerobic Oxidative Radical Addition ...
    Dec 27, 2022 · We reported the breakthrough in diastereo- and enantioselective catalytic aerobic oxidative radical addition/hydroxylation/1,4-aryl migration processes.
  15. [15]
    Green strategies for transition metal-catalyzed C–H activation in ...
    Jun 2, 2021 · Transition metal-catalyzed C–H activation has emerged as viable tool for molecular synthesis by reducing the formation of undesired byproducts.
  16. [16]
    C–H Activation: Toward Sustainability and Applications
    Feb 2, 2021 · While internal and organic oxidants avoid the use of transition metals, these strategies are still responsible for stoichiometric waste.
  17. [17]
    Homogeneous Catalysis - an overview | ScienceDirect Topics
    Most homogeneous catalysis reactions involve two processes known as Oxidative Addition and Reductive Elimination as shown in Fig. 11.1. Sign in to download full ...
  18. [18]
    Uranium-mediated oxidative addition and reductive elimination
    Abstract. Oxidative addition, and its reverse reaction reductive elimination, constitute two key reactions that underpin organometallic chemistry and catalysis.
  19. [19]
    Oxidative Addition and Reductive Elimination - Wiley Online Library
    Mar 28, 2014 · In oxidative addition, the concerted, S N 2, radical and ionic mechanisms are outlined. Reductive elimination goes by the reverse pathways in ...Missing: review | Show results with:review
  20. [20]
    10.1: Oxidative Addition and Reductive Elimination
    Feb 24, 2023 · The oxidative addition is facilitated by electron rich metal centers having low oxidation state whereas the reductive elimination is facilitated ...
  21. [21]
    Organometallic HyperTextBook: Reductive Elimination
    The rates of reductive elimination (as well as oxidative addition) depend on a variety of factors, but one of the most important is thermodynamics. Consider ...
  22. [22]
    Mechanistic Aspects of the Palladium‐Catalyzed Suzuki‐Miyaura ...
    Jul 16, 2021 · The Suzuki-Miyaura reaction uses palladium to form C-C bonds with organoboron compounds and electrophiles, involving oxidative addition, ...
  23. [23]
    Palladium-Catalyzed Intramolecular Reactions of (E)
    A full account of the Pd-catalyzed intramolecular reactions of (E)-2,2-disubstituted 1-alkenyldimethylalanes with aryl triflates as an entry into ...
  24. [24]
    Oxidative Addition - an overview | ScienceDirect Topics
    Oxidative addition is defined as the process by which a molecule AX dissociates its A-X bond to coordinate two fragments to a complex, often occurring in ...<|control11|><|separator|>
  25. [25]
  26. [26]
    Reversible Addition of Sulfur Dioxide to Four-Coordinated Metal ...
    Structure and bonding of transition metal-sulfur dioxide complexes. 1981, 47-100. https://doi.org/10.1007/3-540-10655-3_2. K. K. Pandey, U. C. Agarwala ...
  27. [27]
    Four‐Electron Oxidative Addition of an N=N Double Bond at a ...
    Sep 1, 2023 · Oxidative addition reactions are ubiquitous across organometallic chemistry constituting critical steps in many of the stoichiometric and ...
  28. [28]
    The mechanism of oxidative addition of Pd(0) to Si–H bonds
    Sep 14, 2021 · The oxidative addition of Pd to Si–H bonds is a crucial step in a variety of catalytic applications, and many aspects of this reaction are ...
  29. [29]
  30. [30]
  31. [31]
    Mechanistic Studies into the Oxidative Addition of Co(I) Complexes
    Nov 8, 2019 · (1) The mechanism of oxidative addition is highly dependent on both the metal complex and the substrate involved and can proceed through a ...
  32. [32]
    Kinetics of the Addition of Hydrogen, Oxygen, and Methyl Iodide to ...
    Kinetic study of the oxidative addition of methyl iodide to Vaska's complex in ionic liquids. Journal of Organometallic Chemistry 2005, 690 (15) , 3522-3528 ...Missing: original | Show results with:original<|control11|><|separator|>
  33. [33]
    [PDF] Kinetic Isotope Effects in the Study of Organometallic Reaction ...
    May 5, 2011 · (203) KIEs in the addition of H2 to Vaska's complex 303 were among the earlier values determined for an organometallic transforma- tion: (a) ...
  34. [34]
    Machine learning dihydrogen activation in the chemical space ...
    Apr 7, 2020 · In this work, we provided an efficient protocol combining DFT and ML calculations for the prediction of reactivity in the oxidative addition of ...
  35. [35]
    Understanding oxidative addition in organometallics: a closer look
    The goal of this study is to present oxidative addition by highlighting the history, fundamental concept, mechanism and scope. Relevance to current applications ...
  36. [36]
  37. [37]
    Oxidative addition, carbonyl insertion, and o-demethylation ...
    The complex trans-[RhCl(CO){PMe2(2-MeOC6H4)}2] undergoes rapid oxidative addition with HCl, MeCl, MeI, CCl4, Cl2, and PhCOCl (R–X) to give rhodium(III) ...
  38. [38]
    Oxidative Addition of Water, Alcohols, and Amines in Palladium ...
    Aug 3, 2020 · The oxidative addition of OH, NH, and weak CH bonds to a palladium(0) complex is facile, allows for additive-free catalysis, and suggests the non-innocence of ...
  39. [39]
    Activation of water molecule. 2. Generation of strong hydroxo bases ...
    A cooperative pathway for water activation using a bimetallic Pt 0 –Cu I system. ... Yoshida, T. Matsuda, S. Otsuka, G. W. Parshall, W. G. Peet, P. Binger ...
  40. [40]
    Bimolecular homolytic displacement of transition-metal complexes ...
    Article September 1, 1983. Bimolecular homolytic displacement of transition-metal complexes from carbon. Click to copy article linkArticle link copied ...
  41. [41]
    [PDF] We Already Know Everything about Oxidative Addition to Pd(0) - HAL
    Nov 14, 2023 · Abstract The oxidative addition (OA) of organic electrophiles to Pd(0) is a fundamental step in organopalladium chemistry and.
  42. [42]
    A radical approach to the copper oxidative addition problem - Science
    Jun 1, 2018 · Here, we demonstrate that this copper oxidative addition problem can be overcome with an aryl radical–capture mechanism, wherein the aryl ...<|separator|>
  43. [43]
    Direct observation of MeRh(CO)2I3-, the key intermediate in ...
    The oxidative addition and migratory 1,1-insertion in the Monsanto and Cativa processes. A density functional study of the catalytic carbonylation of methanol.
  44. [44]
    Study on Rh(I)/Ru(III) Bimetallic Catalyst Catalyzed Carbonylation of ...
    Sep 11, 2020 · In the Monsanto rhodium/iodine catalyst catalyzed carbonylation to acetic acid reaction, Forster [8] proposed that the reaction mainly ...
  45. [45]
    Monsanto (BP) Acetic Acid & Related Processes - Academia.edu
    The negative charge on the [Rh(CO)2I2] catalyst is believed to be critical in assisting the oxidative addition of MeI to the rhodium center. The alkyl species, ...
  46. [46]
    Economic assessment of acetaldehyde production from oxidative ...
    The primary industrial process for acetaldehyde production is the Wacker process [5], which converts fossil-based ethylene to acetaldehyde using a PdCl2/CuCl2 ...
  47. [47]
    Wacker Chemie - an overview | ScienceDirect Topics
    Later on, a one-stage process was developed by Farbwerke Hoechst that employed catalytic amount of CuCl2 (23). The key to success of the Wacker process is ...
  48. [48]
    The mechanism of the Wacker process. Corroborative evidence for ...
    The mechanism of the Wacker process. Corroborative evidence for distal addition of water and palladium | Organometallics. Pure Hydrocarbons from Petroleum.
  49. [49]
    Applied Hydroformylation | Chemical Reviews - ACS Publications
    Hydroformylation is the addition of synthesis gas (“syngas”), a mixture of CO and H2, to olefins in the presence of a catalyst under the formation of ...
  50. [50]
    A retrospective-prospective review of Suzuki–Miyaura reaction
    Nov 15, 2022 · The Suzuki coupling reaction can produce some drugs like losartan, valsartan, diflunisal, lapatinib, atazanavir, and clonazepam.
  51. [51]
    Nuances in Fundamental Suzuki–Miyaura Cross-Couplings ...
    May 23, 2018 · We have explored fundamental Pd-catalyzed C sp 2 –C sp 2 Suzuki–Miyaura cross-couplings of aryl iodides (Ar–I) employing “classical” Pd/PPh 3 catalyst systems.
  52. [52]
    Recycling of Homogeneous Catalysts Basic Principles, Industrial ...
    Apr 25, 2024 · One drawback is the potential incorporation of the catalyst into the precipitated material, necessitating additional purification steps.
  53. [53]
    Recovery of Homogeneous Platinoid Catalysts from Pharmaceutical ...
    Catalyst recovery is a major challenge for reaching the objectives of green chemistry for industry. Indeed, catalysts enable quick and selective syntheses ...
  54. [54]
    Understanding oxidative addition in organometallics: a closer look
    Dec 2, 2021 · The goal of this study is to present oxidative addition by highlighting the history, fundamental concept, mechanism and scope.
  55. [55]
    Single-Electron-Transfer-Mediated Carbonylation Reactions
    Mar 5, 2025 · This strategy focuses on designing novel SET-mediated carbonylation reactions for unactivated alkyl halides, intending to avoid harsh reaction ...
  56. [56]
    Tracking C–H activation with orbital resolution - Science
    Jun 1, 2023 · Svensson, Different electronic structure requirements on precursors and transition states for the oxidative addition reaction with methane.
  57. [57]
    AI Approaches to Homogeneous Catalysis with Transition Metal ...
    May 14, 2025 · Two distinct ML models were trained with this data to predict the oxidative addition barriers: neural networks (NN) and GPs. The NN models were ...
  58. [58]
    Defying the oxidative-addition prerequisite in cross-coupling through ...
    Apr 4, 2025 · The transition metal-catalyzed cross-coupling reaction is a cornerstone for organic synthesis, playing a crucial role in the fine chemical and ...
  59. [59]
    On-Demand Access to Palladium Oxidative Addition Complexes ...
    Feb 28, 2025 · We report a simple admixture approach combining a bench-stable organopalladate salt with a mono- or bidentate ligand to generate OACs within minutes at room ...
  60. [60]
  61. [61]
    Pd-catalyzed asymmetric P–C cross-coupling to access ortho-NH2 or
    Aug 29, 2025 · It is noteworthy that the amino-phosphine oxide product tended to recrystallize, which improved the ee value of 3a to 99%. The absolute ...