Wilkinson's catalyst

From Wikipedia, the free encyclopedia

Wilkinson's catalyst
Wilkinson's catalyst
Wilkinson's catalyst
Wilkinson's catalyst
Wilkinson's
Names
IUPAC name
(SP-4)-chlorido­tris(triphenylphosphene)­rhodium
Other names
Rhodium(I) tris(triphenylphosphene) chloride,
Wilkinson's catalyst,
Tris(triphenylphosphene)­rhodium(I) chloride
Identifiers
3D model (JSmol)
ECHA InfoCard 100.035.207 Edit this at Wikidata
EC Number
  • 238-744-5
RTECS number
  • none
UNII
  • InChI=1S/3C18H15P.ClH.Rh/c3*1-4-10-16(11-5-1)19(17-12-6-2-7-13-17)18-14-8-3-9-15-18;;/h3*1-15H;1H;/p-1
    Key: QBERHIJABFXGRZ-UHFFFAOYSA-M
  • Cl[Rh-3]([P+](c0ccccc0)(c0ccccc0)c0ccccc0)([P+](c0ccccc0)(c0ccccc0)c0ccccc0)[P+](c0ccccc0)(c0ccccc0)c0ccccc0
Properties
C54H45ClP3Rh
Molar mass 925.22 g/mol
Appearance red solid
Melting point 245 to 250 °C (473 to 482 °F; 518 to 523 K)
insoluble in water
Solubility in other solvents 20 g/L (CHCl3, CH2Cl2), 2 g/L (benzene, toluene)[1]
Structure
square planar d8 (diamagnetic; sp2d-hybridized)
Hazards[2]
Occupational safety and health (OHS/OSH):
Main hazards
none
GHS labelling:
GHS07: Exclamation mark
Warning
H302, H317, H413
P261, P264, P270, P272, P273, P280, P301+P312, P302+P352, P330, P333+P313, P363, P501
Related compounds
Related compounds
triphenylphosphene
Pd(PPh3)4
IrCl(CO)[P(C6H5)3]2
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
☒N verify (what is checkY☒N ?)

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.

Historically, Wilkinson's catalyst has been a paradigm in catalytic studies leading to several advances in the field such as the implementation of some of the first heteronuclear magnetic resonance studies for its structural elucidation in solution (31P),[3] parahydrogen-induced polarization spectroscopy to determine the nature of transient reactive species,[4] or one of the first detailed kinetic investigation by Halpern to elucidate the mechanism.[5] Furthermore, the catalytic and organometallic studies on Wilkinson's catalyst also played a significant role on the subsequent development of cationic Rh- and Ru-based asymmetric hydrogenation transfer catalysts which set the foundations for modern asymmetric catalysis.[6]

Structure and basic properties[edit]

According to single crystal X-ray diffraction the compound adopts a slightly distorted square planar structure.[7]

In analyzing the bonding, it is a complex of Rh(I), a d8 transition metal ion. From the perspective of the 18-electron rule, the four ligands each provides two electrons, for a total of 16-electrons. As such the compound is coordinatively unsaturated, i.e. susceptible to binding substrates (alkenes and H2). In contrast, IrCl(PPh3)3 undergoes cyclometallation to give HIrCl(PPh3)2(PPh2C6H4), a coordinatively saturated Ir(III) complex that is not catalytically active.[8]

Synthesis[edit]

Wilkinson's catalyst is usually obtained by treating rhodium(III) chloride hydrate with an excess of triphenylphosphine in refluxing ethanol.[9][10][1] Triphenylphosphine serves as both a ligand and a two-electron reducing agent that oxidizes itself from oxidation state (III) to (V). In the synthesis, three equivalents of triphenylphosphine become ligands in the product, while the fourth reduces rhodium(III) to rhodium(I).

RhCl3(H2O)3 + 4 PPh3 → RhCl(PPh3)3 + OPPh3 + 2 HCl + 2 H2O

Catalytic applications[edit]

Wilkinson's catalyst is best known for catalyzing the hydrogenation of olefins with molecular hydrogen.[11][12] The mechanism of this reaction involves the initial dissociation of one or two triphenylphosphine ligands to give 14- or 12-electron complexes, respectively, followed by oxidative addition of H2 to the metal. Subsequent π-complexation of alkene, migratory insertion (intramolecular hydride transfer or olefin insertion), and reductive elimination complete the formation of the alkane product, e.g.:[13]

In terms of their rates of hydrogenation, the degree of substitution on the olefin substrate is the key factor, since the rate-limiting step in the mechanism is the insertion into the olefin which is limited by the severe steric hindrance around the metal center. In practice, terminal and disubstituted alkenes are good substrates, but more hindered alkenes are slower to hydrogenate. The hydrogenation of alkynes is troublesome to control since alkynes tend to be reduced to alkanes, via intermediacy of the cis-alkene.[14] Ethylene reacts with Wilkinson's catalyst to give RhCl(C2H4)(PPh3)2, but it is not a substrate for hydrogenation.[10]

Related catalytic processes[edit]

Wilkinson's catalyst also catalyzes many other hydrofunctionalization reactions including hydroacylation, hydroboration, and hydrosilylation of alkenes.[14] Hydroborations have been studied with catecholborane and pinacolborane.[15] It is also active for the hydrosilylation of alkenes.[16]

In the presence of strong base and hydrogen, Wilkinson's catalyst forms reactive Rh(I) species with superior catalytic activities on the hydrogenation of internal alkynes and functionalized tri-substituted alkenes.[17]

Reactions[edit]

RhCl(PPh3)3 reacts with carbon monoxide to give bis(triphenylphosphine)rhodium carbonyl chloride, trans-RhCl(CO)(PPh3)2. The same complex arises from the decarbonylation of aldehydes:

RhCl(PPh3)3 + RCHO → RhCl(CO)(PPh3)2 + RH + PPh3

Upon stirring in benzene solution, RhCl(PPh3)3 converts to the poorly soluble red-colored dimer [RhCl(PPh3)2]2. This conversion further demonstrates the lability of the triphenylphosphine ligands.

In the presence of base, H2, and additional triphenylphosphine, Wilkinson's complex converts to hydridotetrakis(triphenylphosphine)rhodium(I), HRh(PPh3)4. This 18e complex is also an active hydrogenation catalyst.[18]

See also[edit]

References[edit]

  1. ^ a b Osborn, J. A.; Wilkinson, G. (1967). "Tris(triphenylphosphine)halorhodium(I)". Inorganic Syntheses. Vol. 10. p. 67. doi:10.1002/9780470132418.ch12. ISBN 9780470132418.
  2. ^ "Chlorotris(triphenylphosphine)rhodium(I)". pubchem.ncbi.nlm.nih.gov.
  3. ^ Meakin, P.; Jesson, J. P.; Tolman, C. A. (1 May 1972). "Nature of chlorotris(triphenylphosphene)rhodium in solution and its reaction with hydrogen". Journal of the American Chemical Society. 94 (9): 3240–3242. doi:10.1021/ja00764a061. ISSN 0002-7863.
  4. ^ Duckett, Simon B.; Newell, Connie L.; Eisenberg, Richard (1994). "Observation of New Intermediates in Hydrogenation Catalyzed by Wilkinson's Catalyst, RhCl(PPh3)3, Using Parahydrogen-Induced Polarization". Journal of the American Chemical Society. 116 (23): 10548–10556. doi:10.1021/ja00102a023.
  5. ^ Halpern, Jack (1 January 1981). "Mechanistic aspects of homogeneous catalytic hydrogenation and related processes". Inorganica Chimica Acta. 50: 11–19. doi:10.1016/S0020-1693(00)83716-0.
  6. ^ Hartwig, John F. (2010). Organotransition metal chemistry- From bonding to Catalysis. University Science Books. ISBN 978-1-891389-53-5.
  7. ^ Bennett, Michael J.; Donaldson, Peter B. (1977). "Crystal and molecular structure of the orange and red allotropes of chlorotris(triphenylphosphine)rhodium(I)". Inorganic Chemistry. 16 (3): 655–660. doi:10.1021/ic50169a033.
  8. ^ Bennett, M. A.; Milner, D. L. (1969). "Chlorotris(triphenylphosphine)iridium(I) and Related Complexes. Oxidative Addition Reactions and Hydrogen Abstraction from the Coordinated Ligand". Journal of the American Chemical Society. 91 (25): 6983–6994. doi:10.1021/ja01053a016.
  9. ^ Bennett, M. A.; Longstaff, P. A. (1965). "Complexes of Rhodium(I) with Triphenylphosphine". Chem. Ind. (London). 1965: 846.
  10. ^ a b Osborn, J. A.; Jardine, F. H.; Young, J. F.; Geoffrey Wilkinson (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". Journal of the Chemical Society A. 1966: 1711–1732. doi:10.1039/J19660001711.
  11. ^ Arthur Birch; Williamson, D. H. (1976). "Homogeneous Hydrogenation Catalysts in Organic Synthesis". Organic Reactions. 24: 1.
  12. ^ James, B. R. (1973). Homogeneous Hydrogenation. New York, NY: John Wiley & Sons.
  13. ^ Rzepa, Henry (21 January 2024). "A mechanistic exploration of the Wilkinson hydrogenation catalyst. Part 1: Model templates". Chemistry with a Twist.
  14. ^ a b Kevin Burgess, Wilfred van der Donk, Chul-Ho Jun, Young Jun Park, "Chlorotris(triphenylphosphine)-rhodium(I)" Encyclopedia of Reagents for Organic Synthesis 2005 John Wiley & Sons. doi:10.1002/047084289X.rc162s.pub2
  15. ^ Evans, D. A.; Fu, G. C.; Hoveyda, A. H. (1988). "Rhodium(I)-catalyzed hydroboration of olefins. The documentation of regio- and stereochemical control in cyclic and acyclic systems". J. Am. Chem. Soc. 110 (20): 6917–6918. doi:10.1021/ja00228a068.
  16. ^ Ojima, I.; Kogure, T. (1972). "Selective reduction of α,β-unsaturated terpene carbonyl compounds using hydrosilane-rhodium(I) complex combinations". Tetrahedron Lett. 13 (49): 5035–5038. doi:10.1016/S0040-4039(01)85162-5.
  17. ^ Perea Buceta, Jesus E.; Fernández, Israel; Heikkinen, Sami; Axenov, Kirill; King, Alistair W. T.; Niemi, Teemu; Nieger, Martin; Leskelä, Markku; Repo, Timo (23 November 2015) [2015]. "Diverting Hydrogenations with Wilkinson's Catalyst towards Highly Reactive Rhodium(I) Species". Angewandte Chemie International Edition. 54 (48): 14321–14325. doi:10.1002/anie.201506216. ISSN 1521-3773. PMID 26437764.
  18. ^ Eduardo Peña-Cabrera "Hydridotetrakis(triphenylphosphine)rhodium" Encyclopedia of Reagents for Organic Synthesis, 2001, John Wiley & Sons. doi:10.1002/047084289X.rh030m