Crabtree's catalyst

Last updated
Crabtree's catalyst
Crabtree.svg
Crabtree's-catalyst-cation-3D-sticks.png
Names
IUPAC name
(SP-4)-(η22-Cycloocta-1,5-diene)(pyridine)(tricyclohexylphosphane)iridium(1+) hexafluoridophosphate(1−)
Identifiers
3D model (JSmol)
ChemSpider
ECHA InfoCard 100.164.161 OOjs UI icon edit-ltr-progressive.svg
PubChem CID
UNII
  • InChI=1S/C18H33P.C8H12.C5H5N.F6P.Ir/c1-4-10-16(11-5-1)19(17-12-6-2-7-13-17)18-14-8-3-9-15-18;1-2-4-6-8-7-5-3-1;1-2-4-6-5-3-1;1-7(2,3,4,5)6;/h16-18H,1-15H2;1-2,7-8H,3-6H2;1-5H;;/q;;;-1;+1/b;2-1-,8-7-;;; X mark.svgN
    Key: UJXHUUQZACSUOG-KJWGIZLLSA-N X mark.svgN
  • InChI=1/C18H33P.C8H12.C5H5N.F6P.Ir/c1-4-10-16(11-5-1)19(17-12-6-2-7-13-17)18-14-8-3-9-15-18;1-2-4-6-8-7-5-3-1;1-2-4-6-5-3-1;1-7(2,3,4,5)6;/h16-18H,1-15H2;1-2,7-8H,3-6H2;1-5H;;/q;;;-1;+1/b;2-1-,8-7-;;;
    Key: UJXHUUQZACSUOG-KJWGIZLLBW
  • c1ccncc1.C1CCC(CC1)P(C2CCCCC2)C3CCCCC3.C1/C=C\CC/C=C\C1.F[P-](F)(F)(F)(F)F.[Ir+]
Properties
C31H50F6IrNP2
Molar mass 804.9026 g/mol
AppearanceYellow microcrystals
Melting point 150 °C (302 °F; 423 K) (decomposes) [1]
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
X mark.svgN  verify  (what is  Yes check.svgYX mark.svgN ?)

Crabtree's catalyst is an organoiridium compound with the formula [ C8H12IrP(C6H11)3 C5H5N ]PF6. It is a homogeneous catalyst for hydrogenation and hydrogen-transfer reactions, developed by Robert H. Crabtree. This air stable orange solid is commercially available and known for its directed hydrogenation to give trans stereoselectivity with respective of directing group. [2] [3]

Contents

Structure and synthesis

The complex has a square planar molecular geometry, as expected for a d8 complex. It is prepared from cyclooctadiene iridium chloride dimer. [4]

Reactivity

Crabtree’s catalyst is effective for the hydrogenations of mono-, di-, tri-, and tetra-substituted substrates. Whereas Wilkinson’s catalyst and the Schrock–Osborn catalyst do not catalyze the hydrogenation of a tetrasubstituted olefin, Crabtree’s catalyst does so to at high turnover frequencies (table). [2] [5]

Turnover frequencies
SubstrateWilkinson's catalystSchrock–Osborn catalystCrabtree's catalyst
Hex-1-ene65040006400
Cyclohexene700104500
1-Methylcyclohexene133800
2,3-Dimethyl-but-2-ene4000

The catalyst is reactive at room temperature. [1] The reaction is robust without drying solvents or meticulous deoxygenation of the hydrogen. The catalyst is tolerant of weakly basic functional groups such as ester, but not alcohols (see below) or amines. [2] The catalyst is sensitive to proton-bearing impurities. [6]

The catalyst becomes irreversibly deactivated after about ten minutes at room temperature, signaled by appearance of yellow color. One deactivation process involves formation of hydride-bridged dimers. [7] As a consequence, Crabtree's Catalyst is usually used in very low catalyst loading.

Crabtree's catalyst is thought to operate via an intermediate such as this: cis-[IrH2(cod)L2] (cationic charge not shown). CrabtreeHydrogen.png
Crabtree's catalyst is thought to operate via an intermediate such as this: cis-[IrH2(cod)L2] (cationic charge not shown).

Other catalytic functions: isotope exchange and isomerization

Besides hydrogenation, the catalyst catalyzes the isomerization and hydroboration of alkenes. [1]

An example of isomerization with Crabtree's catalyst. The reaction proceeds 98% to completion in 30 minutes at room temperature. CrabtreeIsomerization1.png
An example of isomerization with Crabtree's catalyst. The reaction proceeds 98% to completion in 30 minutes at room temperature.

Crabtree's catalyst is used in isotope exchange reactions. More specifically, it catalyzes the direct exchange of a hydrogen atom with its isotopes deuterium and tritium, without the use of an intermediate. [8] It has been shown that isotope exchange with Crabtree’s catalyst is highly regioselective. [9] [10]

Influence of directing functional groups

The hydrogenation of a terpen-4-ol demonstrates the ability of compounds with directing groups (the –OH group) to undergo diastereoselective hydrogenation. With palladium on carbon in ethanol the product distribution is 20:80 favoring the cis isomer (2B in Scheme 1). The polar side (with the hydroxyl group) interacts with the solvent. This is due to slight haptophilicity, an effect in which a functional group binds to the surface of a heterogeneous catalyst and directs the reaction. [11] [12] In cyclohexane as solvent, the distribution changes to 53:47 because haptophilicity is no long present (there is no directing group on cyclohexane). The distribution changes completely in favor of the trans isomer 2A when Crabtree's catalyst is used in dichloromethane. This selectivity is both predictable and practically useful. [13] Carbonyl groups are also known to direct the hydrogenation by the Crabtree catalyst to be highly regioselective. [14] [15] [16]

Crabtree cat reaction.png

The directing effect that causes the stereoselectivity of hydrogenation of terpen-4-ol with Crabtree’s catalyst is shown below.

Directing effect of an -OH group on diastereoselectivity of hydrogenation by Crabtree's catalyst. Hydrogen is added from the direction of the iridium atom, selecting for the reactivity shown above. Additional ligands on catalyst not shown. CrabtreeDirectingEffect.png
Directing effect of an –OH group on diastereoselectivity of hydrogenation by Crabtree's catalyst. Hydrogen is added from the direction of the iridium atom, selecting for the reactivity shown above. Additional ligands on catalyst not shown.

History

Crabtree and graduate student George Morris discovered this catalyst in the 1970s while working on iridium analogues of Wilkinson's rhodium-based catalyst at the Institut de Chimie des Substances Naturelles at Gif-sur-Yvette, near Paris.

Previous hydrogenation catalysts included Wilkinson’s catalyst and a cationic rhodium(I) complex with two phosphine groups developed by Osborn and Schrock. [17] These catalysts accomplished hydrogenation through displacement; after hydrogen addition across the metal, a solvent or a phosphine group dissociated from the rhodium metal so the olefin to be hydrogenated could gain access to the active site. [2] This displacement occurs quickly for rhodium complexes but occurs barely at all for iridium complexes. [18] Because of this, research at the time focused on rhodium compounds instead of compounds involving transition metals of the third row, like iridium. Wilkinson, Osborn, and Schrock also only used coordinating solvents. [19]

Crabtree noted that the ligand dissociation step does not occur in heterogeneous catalysis, and so posited that this step was limiting in homogeneous systems. [2] They sought catalysts with "irreversibly created active sites in a noncoordinating solvent." This led to the development of the Crabtree catalyst, and use of the solvent CH2Cl2.

Related Research Articles

<span class="mw-page-title-main">Organometallic chemistry</span> Study of organic compounds containing metal(s)

Organometallic chemistry is the study of organometallic compounds, chemical compounds containing at least one chemical bond between a carbon atom of an organic molecule and a metal, including alkali, alkaline earth, and transition metals, and sometimes broadened to include metalloids like boron, silicon, and selenium, as well. Aside from bonds to organyl fragments or molecules, bonds to 'inorganic' carbon, like carbon monoxide, cyanide, or carbide, are generally considered to be organometallic as well. Some related compounds such as transition metal hydrides and metal phosphine complexes are often included in discussions of organometallic compounds, though strictly speaking, they are not necessarily organometallic. The related but distinct term "metalorganic compound" refers to metal-containing compounds lacking direct metal-carbon bonds but which contain organic ligands. Metal β-diketonates, alkoxides, dialkylamides, and metal phosphine complexes are representative members of this class. The field of organometallic chemistry combines aspects of traditional inorganic and organic chemistry.

<span class="mw-page-title-main">Hydrogenation</span> Chemical reaction between molecular hydrogen and another compound or element

Hydrogenation is a chemical reaction between molecular hydrogen (H2) and another compound or element, usually in the presence of a catalyst such as nickel, palladium or platinum. The process is commonly employed to reduce or saturate organic compounds. Hydrogenation typically constitutes the addition of pairs of hydrogen atoms to a molecule, often an alkene. Catalysts are required for the reaction to be usable; non-catalytic hydrogenation takes place only at very high temperatures. Hydrogenation reduces double and triple bonds in hydrocarbons.

In organic chemistry, hydroformylation, also known as oxo synthesis or oxo process, is an industrial process for the production of aldehydes from alkenes. This chemical reaction entails the net addition of a formyl group and a hydrogen atom to a carbon-carbon double bond. This process has undergone continuous growth since its invention: production capacity reached 6.6×106 tons in 1995. It is important because aldehydes are easily converted into many secondary products. For example, the resultant aldehydes are hydrogenated to alcohols that are converted to detergents. Hydroformylation is also used in speciality chemicals, relevant to the organic synthesis of fragrances and pharmaceuticals. The development of hydroformylation is one of the premier achievements of 20th-century industrial chemistry.

In chemistry, homogeneous catalysis is catalysis where the catalyst is in same phase as reactants, principally by a soluble catalyst a in solution. In contrast, heterogeneous catalysis describes processes where the catalysts and substrate are in distinct phases, typically solid-gas, respectively. The term is used almost exclusively to describe solutions and implies catalysis by organometallic compounds. Homogeneous catalysis is an established technology that continues to evolve. An illustrative major application is the production of acetic acid. Enzymes are examples of homogeneous catalysts.

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

Wilkinson's catalyst (chlorido­tris(triphenylphosphene)­rhodium(I)) is a coordination complex of rhodium with the formula [RhCl(PPh3)3], where 'Ph' denotes a phenyl group. 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.

<span class="mw-page-title-main">Olefin metathesis</span> Organic reaction involving the breakup and reassembly of alkene double bonds

In organic chemistry, olefin metathesis is an organic reaction that entails the redistribution of fragments of alkenes (olefins) by the scission and regeneration of carbon-carbon double bonds. Because of the relative simplicity of olefin metathesis, it often creates fewer undesired by-products and hazardous wastes than alternative organic reactions. For their elucidation of the reaction mechanism and their discovery of a variety of highly active catalysts, Yves Chauvin, Robert H. Grubbs, and Richard R. Schrock were collectively awarded the 2005 Nobel Prize in Chemistry.

<span class="mw-page-title-main">Rhodium(III) chloride</span> Chemical compound

Rhodium(III) chloride refers to inorganic compounds with the formula RhCl3(H2O)n, where n varies from 0 to 3. These are diamagnetic solids featuring octahedral Rh(III) centres. Depending on the value of n, the material is either a dense brown solid or a soluble reddish salt. The soluble trihydrated (n = 3) salt is widely used to prepare compounds used in homogeneous catalysis, notably for the industrial production of acetic acid and hydroformylation.

<span class="mw-page-title-main">Bamford–Stevens reaction</span> Synthesis of alkenes by base-catalysed decomposition of tosylhydrazones

The Bamford–Stevens reaction is a chemical reaction whereby treatment of tosylhydrazones with strong base gives alkenes. It is named for the British chemist William Randall Bamford and the Scottish chemist Thomas Stevens Stevens (1900–2000). The usage of aprotic solvents gives predominantly Z-alkenes, while protic solvent gives a mixture of E- and Z-alkenes. As an alkene-generating transformation, the Bamford–Stevens reaction has broad utility in synthetic methodology and complex molecule synthesis.

<span class="mw-page-title-main">Robert H. Crabtree</span> British-American chemist

Robert Howard Crabtree is a British-American chemist. He is serving as Conkey P. Whitehead Professor Emeritus of Chemistry at Yale University in the United States. He is a naturalized citizen of the United States. Crabtree is particularly known for his work on "Crabtree's catalyst" for hydrogenations, and his textbook on organometallic chemistry.

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".

Hydrosilylation, also called catalytic hydrosilation, describes the addition of Si-H bonds across unsaturated bonds. Ordinarily the reaction is conducted catalytically and usually the substrates are unsaturated organic compounds. Alkenes and alkynes give alkyl and vinyl silanes; aldehydes and ketones give silyl ethers. Hydrosilylation has been called the "most important application of platinum in homogeneous catalysis."

Martin Arthur Bennett FRS is an Australian inorganic chemist. He gained recognition for studies on the co-ordination chemistry of tertiary phosphines, olefins, and acetylenes, and the relationship of their behaviour to homogeneous catalysis.

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">Organoiridium chemistry</span> Chemistry of organometallic compounds containing an iridium-carbon bond

Organoiridium chemistry is the chemistry of organometallic compounds containing an iridium-carbon chemical bond. Organoiridium compounds are relevant to many important processes including olefin hydrogenation and the industrial synthesis of acetic acid. They are also of great academic interest because of the diversity of the reactions and their relevance to the synthesis of fine chemicals.

Asymmetric hydrogenation is a chemical reaction that adds two atoms of hydrogen to a target (substrate) molecule with three-dimensional spatial selectivity. Critically, this selectivity does not come from the target molecule itself, but from other reagents or catalysts present in the reaction. This allows spatial information to transfer from one molecule to the target, forming the product as a single enantiomer. The chiral information is most commonly contained in a catalyst and, in this case, the information in a single molecule of catalyst may be transferred to many substrate molecules, amplifying the amount of chiral information present. Similar processes occur in nature, where a chiral molecule like an enzyme can catalyse the introduction of a chiral centre to give a product as a single enantiomer, such as amino acids, that a cell needs to function. By imitating this process, chemists can generate many novel synthetic molecules that interact with biological systems in specific ways, leading to new pharmaceutical agents and agrochemicals. The importance of asymmetric hydrogenation in both academia and industry contributed to two of its pioneers — William Standish Knowles and Ryōji Noyori — being collectively awarded one half of the 2001 Nobel Prize in Chemistry.

<span class="mw-page-title-main">Organorhodium chemistry</span> Field of study

Organorhodium chemistry is the chemistry of organometallic compounds containing a rhodium-carbon chemical bond, and the study of rhodium and rhodium compounds as catalysts in organic reactions.

<span class="mw-page-title-main">Peter Maitlis</span> British chemist (1933–2022)

Peter Michael Maitlis, FRS was a British organometallic chemist.

Dehydrogenation of amine-boranes or dehydrocoupling of amine-boranes is a chemical process in main group and organometallic chemistry wherein dihydrogen is released by the coupling of two or more amine-borane adducts. This process is of due to the potential of using amine-boranes for hydrogen storage.

<span class="mw-page-title-main">John A. Osborn</span> English inorganic chemist

John A. Osborn (1939–2000) was an inorganic chemist who made many contributions to organometallic chemistry. Obsorn received his PhD under the mentorship of Geoffrey Wilkinson. During that degree Osborn contributed to the development of Wilkinson's catalyst. His thesis studies ranged widely.

Iridium compounds are compounds containing the element iridium (Ir). Iridium forms compounds in oxidation states between −3 and +9, but the most common oxidation states are +1, +2, +3, and +4. Well-characterized compounds containing iridium in the +6 oxidation state include IrF6 and the oxides Sr2MgIrO6 and Sr2CaIrO6. iridium(VIII) oxide was generated under matrix isolation conditions at 6 K in argon. The highest oxidation state (+9), which is also the highest recorded for any element, is found in gaseous [IrO4]+.

References

  1. 1 2 3 Crabtree, R. H. (2001). "(1,5-Cyclooctadiene)(tricyclohexylphosphine)(pyridine)iridium(I) Hexafluorophosphate". e-EROS Encyclopedia of Reagents for Organic Synthesis. doi:10.1002/047084289X.rc290m.pub4 (inactive 31 January 2024).{{cite encyclopedia}}: CS1 maint: DOI inactive as of January 2024 (link)
  2. 1 2 3 4 5 Robert H. Crabtree (1979). "Iridium compounds in catalysis". Acc. Chem. Res. 12 (9): 331–337. doi:10.1021/ar50141a005.
  3. Brown, J. M. (1987). "Directed Homogeneous Hydrogenation". Angew. Chem. Int. Ed. 26 (3): 190–203. doi:10.1002/anie.198701901.
  4. Crabtree, R. H.; Morris, G. E. (1977). "Some diolefin complexes of Iridium(I) and a trans-Influence Series for the Complexes [IrCl(cod)L]". J. Organomet. Chem. 135 (3): 395–403. doi:10.1016/S0022-328X(00)88091-2.
  5. White, M. (2002-10-15). "Hydrogenation" (PDF). Retrieved 2014-12-01.
  6. Xu, Yingjian; Mingos, D. Michael P.; Brown, John M. (2008). "Crabtree's catalyst revisited; Ligand effects on stability and durability". Chem. Comm. 2008 (2): 199–201. doi:10.1039/b711979h. PMID   18092086.
  7. Crabtree, R.; Felkin, H.; Morris, G. (1977). "Cationic iridium diolefin complexes as alkene hydrogenation catalysts and the isolation of some related hydrido complexes". Journal of Organometallic Chemistry . 141 (2): 205–215. doi:10.1016/S0022-328X(00)92273-3.
  8. Schou, S. (2009). "The effect of adding Crabtree's catalyst to rhodium black in direct hydrogen isotope exchange reactions". Journal of Labelled Compounds and Radiopharmaceuticals. 52 (9): 376–381. doi:10.1002/jlcr.1612.
  9. Valsborg, J.; Sorensen, L.; Foged, C. (2001). "Organoiridium catalysed hydrogen isotope exchange of benzamide derivatives". Journal of Labelled Compounds and Radiopharmaceuticals. 44 (3): 209–214. doi:10.1002/jlcr.446.
  10. Hesk, D.; Das, P.; Evans, B. (1995). "Deuteration of acetanilides and other substituted aromatics using [Ir(COD)(Cy3P)(Py)]PF6 as catalyst". Journal of Labelled Compounds and Radiopharmaceuticals. 36 (5): 497–502. doi:10.1002/jlcr.2580360514.
  11. Thompson, H.; Naipawer, R. (1973). "Stereochemical control of reductions. III. Approach to group haptophilicities". Journal of the American Chemical Society . 95 (19): 6379–6386. doi:10.1021/ja00800a036.
  12. Rowlands, G. (2002-01-01). "Hydrogenation" (PDF). Archived from the original (PDF) on 2015-01-02. Retrieved 2014-12-01.
  13. Brown, J. (1987). "Directed Homogeneous Hydrogenation [New Synthetic Methods (65)]". Angew. Chem. Int. Ed. Engl. 26 (3): 190–203. doi:10.1002/anie.198701901.
  14. Schultz, A.; McCloskey, P. (1985). "Carboxamide and carbalkoxy group directed stereoselective iridium-catalyzed homogeneous olefin hydrogenations". Journal of Organic Chemistry . 50 (26): 5905–5907. doi:10.1021/jo00350a105.
  15. Crabtree, R. H.; Davis, M. W. (1986). "Directing effects in homogeneous hydrogenation with [Ir(cod)(PCy3)(py)]PF6". J. Org. Chem. 51 (14): 2655–2661. doi:10.1021/jo00364a007.
  16. Crabtree, R.; Davis, M. (1983). "Occurrence and origin of a pronounced directing effect of a hydroxyl group in hydrogenation with [Ir(cod)P(C6H11)3(py)]PF6". Organometallics. 2 (5): 681–682. doi:10.1021/om00077a019.
  17. Schrock, R.; Osborn, J. A. (1976). "Catalytic hydrogenation using cationic rhodium complexes. I. Evolution of the catalytic system and the hydrogenation of olefins". Journal of the American Chemical Society. 98 (8): 2134–2143. doi:10.1021/ja00424a020.
  18. Osborn, J.; Shapley, J. (1970). "Rapid intramolecular rearrangements in pentacoordinate transition metal compounds. Rearrangement mechanism of some fluxional iridium(I) complexes". Journal of the American Chemical Society. 92 (23): 6976–6978. doi:10.1021/ja00726a047.
  19. Young, J.; Wilkinson, G. (1966). "The preparation and properties of tris(triphenylphosphine)halogenorhodium(I) and some reactions thereof including catalytic homogeneous hydrogenation of olefins and acetylenes and their derivatives". J. Chem. Soc. A. 1966: 1711. doi:10.1039/J19660001711.