Reductive elimination

Last updated

Reductive elimination is an elementary step in organometallic chemistry in which the oxidation state of the metal center decreases while forming a new covalent bond between two ligands. It is the microscopic reverse of oxidative addition, and is often the product-forming step in many catalytic processes. Since oxidative addition and reductive elimination are reverse reactions, the same mechanisms apply for both processes, and the product equilibrium depends on the thermodynamics of both directions. [1] [2]

Contents

General information

Reductive elimination is often seen in higher oxidation states, and can involve a two-electron change at a single metal center (mononuclear) or a one-electron change at each of two metal centers (binuclear, dinuclear, or bimetallic). [1] [2]

General Reductive Elimination.png

For mononuclear reductive elimination, the oxidation state of the metal decreases by two, while the d-electron count of the metal increases by two. This pathway is common for d8 metals Ni(II), Pd(II), and Au(III) and d6 metals Pt(IV), Pd(IV), Ir(III), and Rh(III). Additionally, mononuclear reductive elimination requires that the groups being eliminated must be cis to one another on the metal center. [3]

Cis vs. Trans.png

For binuclear reductive elimination, the oxidation state of each metal decreases by one, while the d-electron count of each metal increases by one. This type of reactivity is generally seen with first row metals, which prefer a one-unit change in oxidation state, but has been observed in both second and third row metals. [4]

Binuclear Reductive Elimination.png

Mechanisms

As with oxidative addition, several mechanisms are possible with reductive elimination. The prominent mechanism is a concerted pathway, meaning that it is a nonpolar, three-centered transition state with retention of stereochemistry. In addition, an SN2 mechanism, which proceeds with inversion of stereochemistry, or a radical mechanism, which proceeds with obliteration of stereochemistry, are other possible pathways for reductive elimination. [1] 3

Octahedral complexes

The rate of reductive elimination is greatly influenced by the geometry of the metal complex. In octahedral complexes, reductive elimination can be very slow from the coordinatively saturated center, and often, reductive elimination only proceeds via a dissociative mechanism, where a ligand must initially dissociate to make a five-coordinate complex. This complex adopts a Y-type distorted trigonal bipyramidal structure where a π-donor ligand is at the basal position and the two groups to be eliminated are brought very close together. After elimination, a T-shaped three-coordinate complex is formed, which will associate with a ligand to form the square planar four-coordinate complex. [5]

Octahedral Reductive Elimination.png

Square planar complexes

Reductive elimination of square planar complexes can progress through a variety of mechanisms: dissociative, nondissociative, and associative. Similar to octahedral complexes, a dissociative mechanism for square planar complexes initiates with loss of a ligand, generating a three-coordinate intermediate that undergoes reductive elimination to produce a one-coordinate metal complex. For a nondissociative pathway, reductive elimination occurs from the four-coordinate system to afford a two-coordinate complex. If the eliminating ligands are trans to each other, the complex must first undergo a trans to cis isomerization before eliminating. In an associative mechanism, a ligand must initially associate with the four-coordinate metal complex to generate a five-coordinate complex that undergoes reductive elimination synonymous to the dissociation mechanism for octahedral complexes. [6] [7]

Reductive Elimination from Square Planar Complexes.png

Factors that affect reductive elimination

Reductive elimination is sensitive to a variety of factors including: 1) metal identity and electron density; 2) sterics; 3) participating ligands; 4) coordination number; 5) geometry; and 6) photolysis/oxidation. Additionally, because reductive elimination and oxidative addition are reverse reactions, any sterics or electronics that enhance the rate of reductive elimination must thermodynamically hinder the rate of oxidative addition. [2]

Metal identity and electron density

First-row metal complexes tend to undergo reductive elimination faster than second-row metal complexes, which tend to be faster than third-row metal complexes. This is due to bond strength, with metal-ligand bonds in first-row complexes being weaker than metal-ligand bonds in third-row complexes. Additionally, electron-poor metal centers undergo reductive elimination faster than electron-rich metal centers since the resulting metal would gain electron density upon reductive elimination. [8]

Electron-Poor Metal Reductive Elimination.png

Sterics

Reductive elimination generally occurs more rapidly from a more sterically hindered metal center because the steric encumbrance is alleviated upon reductive elimination. Additionally, wide ligand bite angles generally accelerate reductive elimination because the sterics force the eliminating groups closer together, which allows for more orbital overlap. [9]

Reductive Elimination Bite Angles.png

Participating ligands

Kinetics for reductive elimination are hard to predict, but reactions that involve hydrides are particularly fast due to effects of orbital overlap in the transition state. [10]

Reductive Elimination Participating Ligands.png

Coordination number

Reductive elimination occurs more rapidly for complexes of three- or five-coordinate metal centers than for four- or six-coordinate metal centers. For even coordination number complexes, reductive elimination leads to an intermediate with a strongly metal-ligand antibonding orbital. When reductive elimination occurs from odd coordination number complexes, the resulting intermediate occupies a nonbonding molecular orbital. [11]

Reductive Elimination Coordination Number.png

Geometry

Reductive elimination generally occurs faster for complexes whose structures resemble the product. [2]

Photolysis/oxidation

Reductive elimination can be induced by oxidizing the metal center to a higher oxidation state via light or an oxidant. [12]

Oxidatively-Induced Reductive Elimination.png

Applications

Reductive elimination has found widespread application in academia and industry, most notable being hydrogenation, [13] the Monsanto acetic acid process, [14] hydroformylation, [15] and cross-coupling reactions. [16] In many of these catalytic cycles, reductive elimination is the product forming step and regenerates the catalyst; however, in the Heck reaction [17] and Wacker process, [18] reductive elimination is involved only in catalyst regeneration, as the products in these reactions are formed via β–hydride elimination.

Related Research Articles

Sharpless asymmetric dihydroxylation is the chemical reaction of an alkene with osmium tetroxide in the presence of a chiral quinine ligand to form a vicinal diol. The reaction has been applied to alkenes of virtually every substitution, often high enantioselectivities are realized, with the chiral outcome controlled by the choice of dihydroquinidine (DHQD) vs dihydroquinine (DHQ) as the ligand. Asymmetric dihydroxylation reactions are also highly site selective, providing products derived from reaction of the most electron-rich double bond in the substrate.

<span class="mw-page-title-main">Organolithium reagent</span> Chemical compounds containing C–Li bonds

In organometallic chemistry, organolithium reagents are chemical compounds that contain carbon–lithium (C–Li) bonds. These reagents are important in organic synthesis, and are frequently used to transfer the organic group or the lithium atom to the substrates in synthetic steps, through nucleophilic addition or simple deprotonation. Organolithium reagents are used in industry as an initiator for anionic polymerization, which leads to the production of various elastomers. They have also been applied in asymmetric synthesis in the pharmaceutical industry. Due to the large difference in electronegativity between the carbon atom and the lithium atom, the C−Li bond is highly ionic. Owing to the polar nature of the C−Li bond, organolithium reagents are good nucleophiles and strong bases. For laboratory organic synthesis, many organolithium reagents are commercially available in solution form. These reagents are highly reactive, and are sometimes pyrophoric.

The Stille reaction is a chemical reaction widely used in organic synthesis. The reaction involves the coupling of two organic groups, one of which is carried as an organotin compound. A variety of organic electrophiles provide the other coupling partner. The Stille reaction is one of many palladium-catalyzed coupling reactions.

Oxidative addition and reductive elimination are two important and related classes of reactions in organometallic chemistry. Oxidative addition is a process that increases both the oxidation state and coordination number of a metal centre. Oxidative addition is often a step in catalytic cycles, in conjunction with its reverse reaction, reductive elimination.

<span class="mw-page-title-main">Wilkinson's catalyst</span> Chemical compound

Wilkinson's catalyst is the common name for chloridotris(triphenylphosphine)rhodium(I), a coordination complex of rhodium with the formula [RhCl(PPh3)3] (Ph = phenyl). It is a red-brown colored solid that is soluble in hydrocarbon solvents such as benzene, and more so in tetrahydrofuran or chlorinated solvents such as dichloromethane. The compound is widely used as a catalyst for hydrogenation of alkenes. It is named after chemist and Nobel laureate Sir Geoffrey Wilkinson, who first popularized its use.

In organic chemistry, carbon–hydrogen bond functionalization is a type of organic reaction in which a carbon–hydrogen bond is cleaved and replaced with a C−X bond. The term usually implies that a transition metal is involved in the C−H cleavage process. Reactions classified by the term typically involve the hydrocarbon first to react with a metal catalyst to create an organometallic complex in which the hydrocarbon is coordinated to the inner-sphere of a metal, either via an intermediate "alkane or arene complex" or as a transition state leading to a "M−C" intermediate. The intermediate of this first step can then undergo subsequent reactions to produce the functionalized product. Important to this definition is the requirement that during the C−H cleavage event, the hydrocarbyl species remains associated in the inner-sphere and under the influence of "M".

In organic chemistry, the Buchwald–Hartwig amination is a chemical reaction for the synthesis of carbon–nitrogen bonds via the palladium-catalyzed coupling reactions of amines with aryl halides. Although Pd-catalyzed C-N couplings were reported as early as 1983, Stephen L. Buchwald and John F. Hartwig have been credited, whose publications between 1994 and the late 2000s established the scope of the transformation. The reaction's synthetic utility stems primarily from the shortcomings of typical methods for the synthesis of aromatic C−N bonds, with most methods suffering from limited substrate scope and functional group tolerance. The development of the Buchwald–Hartwig reaction allowed for the facile synthesis of aryl amines, replacing to an extent harsher methods while significantly expanding the repertoire of possible C−N bond formation.

<span class="mw-page-title-main">Bite angle</span>

In coordination chemistry the bite angle is the ligand–metal–ligand bond angle of coordination complex containing a bidentate ligand. This geometric parameter is used to classify chelating ligands, including those in organometallic complexes. It is most often discussed in terms of catalysis, as changes in bite angle can affect not just the activity and selectivity of a catalytic reaction but even allow alternative reaction pathways to become accessible.

In organometallic chemistry, a migratory insertion is a type of reaction wherein two ligands on a metal complex combine. It is a subset of reactions that very closely resembles the insertion reactions, and both are differentiated by the mechanism that leads to the resulting stereochemistry of the products. However, often the two are used interchangeably because the mechanism is sometimes unknown. Therefore, migratory insertion reactions or insertion reactions, for short, are defined not by the mechanism but by the overall regiochemistry wherein one chemical entity interposes itself into an existing bond of typically a second chemical entity e.g.:

<span class="mw-page-title-main">Organomolybdenum chemistry</span>

Organomolybdenum chemistry is the chemistry of chemical compounds with Mo-C bonds. The heavier group 6 elements molybdenum and tungsten form organometallic compounds similar to those in organochromium chemistry but higher oxidation states tend to be more common.

<span class="mw-page-title-main">Iron tetracarbonyl dihydride</span> Chemical compound

Iron tetracarbonyl dihydride is the organometallic compound with the formula H2Fe(CO)4. This compound was the first transition metal hydride discovered. The complex is stable at low temperatures but decomposes rapidly at temperatures above –20 °C.

An insertion reaction is a chemical reaction where one chemical entity interposes itself into an existing bond of typically a second chemical entity e.g.:

Nucleophilic abstraction is a type of an organometallic reaction which can be defined as a nucleophilic attack on a ligand which causes part or all of the original ligand to be removed from the metal along with the nucleophile.

In chemistry, mesoionic carbenes (MICs) are a type of reactive intermediate that are related to N-heterocyclic carbenes (NHCs); thus, MICs are also referred to as abnormal N-heterocyclic carbenes (aNHCs) or remote N-heterocyclic carbenes (rNHCs). Unlike simple NHCs, the canonical resonance structures of these carbenes are mesoionic: an MIC cannot be drawn without adding additional charges to some of the atoms.

In inorganic chemistry, the cis effect is defined as the labilization of CO ligands that are cis to other ligands. CO is a well-known strong pi-accepting ligand in organometallic chemistry that will labilize in the cis position when adjacent to ligands due to steric and electronic effects. The system most often studied for the cis effect is an octahedral complex M(CO)
5
X
where X is the ligand that will labilize a CO ligand cis to it. Unlike the trans effect, which is most often observed in 4-coordinate square planar complexes, the cis effect is observed in 6-coordinate octahedral transition metal complexes. It has been determined that ligands that are weak sigma donors and non-pi acceptors seem to have the strongest cis-labilizing effects. Therefore, the cis effect has the opposite trend of the trans-effect, which effectively labilizes ligands that are trans to strong pi accepting and sigma donating ligands.

<span class="mw-page-title-main">Transition metal alkyl complexes</span> Coordination complex

Transition metal alkyl complexes are coordination complexes that contain a bond between a transition metal and an alkyl ligand. Such complexes are not only pervasive but are of practical and theoretical interest.

In chemistry, compounds of palladium(III) feature the noble metal palladium in the unusual +3 oxidation state. Compounds of Pd(III) occur in mononuclear and dinuclear forms. Palladium(III) is most often invoked, not observed in mechanistic organometallic chemistry.

In homogeneous catalysis, C2-symmetric ligands refer to ligands that lack mirror symmetry but have C2 symmetry. Such ligands are usually bidentate and are valuable in catalysis. The C2 symmetry of ligands limits the number of possible reaction pathways and thereby increases enantioselectivity, relative to asymmetrical analogues. C2-symmetric ligands are a subset of chiral ligands. Chiral ligands, including C2-symmetric ligands, combine with metals or other groups to form chiral catalysts. These catalysts engage in enantioselective chemical synthesis, in which chirality in the catalyst yields chirality in the reaction product.

Heterobimetallic catalysis is an approach to catalysis that employs two different metals to promote a chemical reaction. Included in this definition are cases where: 1) each metal activates a different substrate, 2) both metals interact with the same substrate, and 3) only one metal directly interacts with the substrate(s), while the second metal interacts with the first.

<span class="mw-page-title-main">Tetrakis(1-norbornyl)cobalt(IV)</span> Chemical compound

Tetrakis(1-norbornyl)cobalt(IV) is an air-sensitive organometallic compound of cobalt. It was first synthesized by Barton K. Bower and Howard G. Tennent in 1972 and is one of few compounds in which cobalt has a formal oxidations state of +4.

References

  1. 1 2 3 Crabtree, Robert H. (2014). The Organometallic Chemistry of the Transition Metals (6 ed.). Wiley. p. 173. ISBN   978-1-118-13807-6.
  2. 1 2 3 4 Hartwig, John F. (2010). Organotransition Metal Chemistry, from Bonding to Catalysis. University Science Books. p. 321. ISBN   978-1-891389-53-5.
  3. Gillie, A.; Stille, J. K. (1980). "Mechanisms of 1,1-Reductive Elimination from Palladium". J. Am. Chem. Soc. 102 (15): 4933–4941. doi:10.1021/ja00535a018.
  4. Okrasinski, S. J.; Nortom, J. R. (1977). "Mechanism of Reductive Elimination. 2. Control of Dinuclear vs. Mononuclear Elimination of Methane from cis-Hydridomethyltetracarbonylosmium". J. Am. Chem. Soc. 99: 295–297. doi:10.1021/ja00443a076.
  5. Milstein, D. (1982). "The First Isolated, Stable cis-Hydridoalkylrhodium Complexes and Their reductive Elimination Reaction". J. Am. Chem. Soc. 104 (19): 5227–5228. doi:10.1021/ja00383a039.
  6. Driver, M. S.; Hartwig, J. F. (1997). "Carbon−Nitrogen-Bond-Forming Reductive Elimination of Arylamines from Palladium(II) Phosphine Complexes". J. Am. Chem. Soc. 119 (35): 8232–8245. doi:10.1021/ja971057x.
  7. Yamamoto, T.; Yamamoto, A.; Ikeda, S. (1971). "Study of Organo(dipyridyl)nickel Complexes. I. Stability and Activation of the Alkyl-Nickel Bonds of Dialkyl(dipyridyl)nickel by Coordination with Various Substituted Olefins". J. Am. Chem. Soc. 93: 3350. doi:10.1021/ja00743a009.
  8. Giovannini, R.; Stüdemann, T.; Dussin, G.; Knochel, P. (1998). "An Efficient Nickel-Catalyzed Cross-Coupling Between sp3 Carbon Centers". Angew. Chem. Int. Ed. 37 (17): 2387–2390. doi:10.1002/(SICI)1521-3773(19980918)37:17<2387::AID-ANIE2387>3.0.CO;2-M. PMID   29710957.
  9. Marcone, J. E.; Moloy, K. G. (1998). "Kinetic Study of Reductive Elimination from the Complexes (Diphosphine)Pd(R)(CN)". J. Am. Chem. Soc. 120 (33): 8527–8528. doi:10.1021/ja980762i.
  10. Low, J. J.; Goddard, III, W. A. (1984). "Reductive Coupling of Hydrogen-Hydrogen, Hydrogen-Carbon, and Carbon-Carbon Bonds from Palladium Complexes". J. Am. Chem. Soc. 106 (26): 8321–8322. doi:10.1021/ja00338a067.
  11. Crumpton-Bregel, D. M.; Goldberg, K. I. (2003). "Mechanisms of C-C and C-H Alkane Reductive Eliminations from Octahedral Pt(IV): Reaction via Five-Coordinate Intermediates or Direct Elimination?". J. Am. Chem. Soc. 125 (31): 9442–9456. doi:10.1021/ja029140u. PMID   12889975.
  12. Lau, W.; Huffman, J. C.; Kochi, J. K. (1982). "Electrochemical Oxidation-Reduction of Organometallic Complexes. Effect of the Oxidation State on the Pathways for Reductive Elimination of Dialkyliron Complexes". Organometallics. 1: 155–169. doi:10.1021/om00061a027.
  13. de Vries, J. G. (2007). The Handbook of Homogeneous Hydrogenation. Wiley. ISBN   978-3-527-31161-3.
  14. Paulik, F. E.; Roth, J. F. (1968). "Novel Catalysts for the Low-pressure Carbonylation of Methanol to Acetic Acid". Chem. Commun. (24): 1578. doi:10.1039/C1968001578A.
  15. Ojima, I.; Tsai, C.-H.; Tzamarioudaki, M.; Bonafoux, D. (2004). "The Hydroformylation Reaction". Organic Reactions. 56: 1–354. doi:10.1002/0471264180.or056.01. ISBN   0471264180.
  16. New Trends in Cross-Coupling: Theory and Applications Thomas Colacot (Editor) 2014 ISBN   978-1-84973-896-5
  17. de Vries, J. G. (2001). "The Heck reaction in the production of fine chemicals" (PDF). Can. J. Chem. 79 (5–6): 1086–1092. doi:10.1139/v01-033. hdl: 11370/31cf3b82-e13a-46e2-a3ba-facb1e2bffbf .
  18. Dong, J. J.; Browne, W. R.; Feringa, B. L. (2015). "Palladium-Catalyzed anti-Markovnikov Oxidation of Terminal Alkenes" (PDF). Angew. Chem. Int. Ed. 54 (3): 734–744. doi:10.1002/anie.201404856. PMID   25367376.