Metal-phosphine complex

Last updated

A metal-phosphine complex is a coordination complex containing one or more phosphine ligands. Almost always, the phosphine is an organophosphine of the type R3P (R = alkyl, aryl). Metal phosphine complexes are useful in homogeneous catalysis. [1] [2] Prominent examples of metal phosphine complexes include Wilkinson's catalyst (Rh(PPh3)3Cl), Grubbs' catalyst, and tetrakis(triphenylphosphine)palladium(0). [3]

Contents

Wilkinson's catalyst, a popular catalyst for hydrogenation. Wilkinson's-catalyst-2D.png
Wilkinson's catalyst, a popular catalyst for hydrogenation.

Preparation

Many metal phosphine complexes are prepared by reactions of metal halides with preformed phosphines. For example, treatment of a suspension of palladium chloride in ethanol with triphenylphosphine yields monomeric bis(triphenylphosphine)palladium(II) chloride units. [4]

[PdCl2]n + 2n PPh3n PdCl2(PPh3)2

The first reported phosphine complexes were cis- and trans-PtCl2(PEt3)2 reported by Cahours and Gal in 1870. [5]

Often the phosphine serves both as a ligand and as a reductant. This property is illustrated by the synthesis of many platinum-metal complexes of triphenylphosphine: [6]

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

M-PR3 bonding

Phosphines are L-type ligands. Unlike most metal ammine complexes, metal phosphine complexes tend to be lipophilic, displaying good solubility in organic solvents.

TEP for selected phosphines [7] (A1 mode of Ni(CO)3L in CH2Cl2)
Lν(CO) cm1
P(t-Bu)32056.1
PMe3 2064.1
PPh3 2068.9
P(OEt)3 2076.3
PCl3 2097.0
PF3 2110.8

Phosphine ligands are also π-acceptors. Their π-acidity arises from overlap of P-C σ* anti-bonding orbitals with filled metal orbitals. Aryl- and fluorophosphines are stronger π-acceptors than alkylphosphines. Trifluorophosphine (PF3) is a strong π-acid with bonding properties akin to those of the carbonyl ligand. [8] In early work, phosphine ligands were thought to utilize 3d orbitals to form M-P pi-bonding, but it is now accepted that d-orbitals on phosphorus are not involved in bonding. [9] The energy of the σ* orbitals is lower for phosphines with electronegative substituents, and for this reason phosphorus trifluoride is a particularly good π-acceptor. [10]

Steric properties

Cone angle is a common and useful parameter for evaluating the steric properties of phosphine ligands. ConeAng.png
Cone angle is a common and useful parameter for evaluating the steric properties of phosphine ligands.

In contrast to tertiary phosphines, tertiary amines, especially arylamine derivatives, are reluctant to bind to metals. The difference between the coordinating power of PR3 and NR3 reflects the greater steric crowding around the nitrogen atom, which is smaller.

By changes in one or more of the three organic substituents, the steric and electronic properties of phosphine ligands can be manipulated. [11] The steric properties of phosphine ligands can be ranked by their Tolman cone angle [7] or percent buried volume. [12]

Spectroscopy

An important technique for the characterization of metal-PR3 complexes is 31P NMR spectroscopy. Substantial shifts occur upon complexation. 31P-31P spin-spin coupling can provide insight into the structure of complexes containing multiple phosphine ligands. [13] [14]

Reactivity

Phosphine ligands are usually "spectator" rather than "actor" ligands. They generally do not participate in reactions, except to dissociate from the metal center. In certain high temperature hydroformylation reactions, the scission of P-C bonds is observed however. [15] The thermal stability of phosphines ligands is enhanced when they are incorporated into pincer complexes.

Applications to homogeneous catalysis

One of the first applications of phosphine ligands in catalysis was the use of triphenylphosphine in "Reppe" chemistry (1948), which included reactions of alkynes, carbon monoxide, and alcohols. [16] In his studies, Reppe discovered that this reaction more efficiently produced acrylic esters using NiBr2(PPh3)2 as a catalyst instead of NiBr2. Shell developed cobalt-based catalysts modified with trialkylphosphine ligands for hydroformylation (now a rhodium catalyst is more commonly used for this process). [17] The success achieved by Reppe and his contemporaries led to many industrial applications. [18]

Illustrative PPh3 complexes

3,3',3''-Phosphanetriyltris(benzenesulfonic acid) trisodium salt forms water-soluble complexes. TPPTS.png
3,3,3-Phosphanetriyltris(benzenesulfonic acid) trisodium salt forms water-soluble complexes.

Complexes of other organophosphorus ligands

The popularity and usefulness of phosphine complexes has led to the popularization of complexes of many related organophosphorus ligands. [5] Complexes of arsines have also been widely investigated, but are avoided in practical applications because of concerns about toxicity.

Complexes of primary and secondary phosphines

Most work focuses on complexes of triorganophosphines, but primary and secondary phosphines, respectively RPH2 and R2PH, also function as ligands. Such ligands are less basic and have small cone angles. These complexes are susceptible to deprotonation leading to phosphido-bridged dimers and oligomers:

2 LnM(PR2H)Cl → [LnM(μ-PR2)]2 + 2 HCl

Complexes of PRx(OR')3−x

Nickel(0) complexes of phosphites, e.g., Ni[P(OEt)3]4 are useful catalysts for hydrocyanation of alkenes. Related complexes are known for phosphinites (R2P(OR')) and phosphonites (RP(OR')2).

Diphosphine complexes

Due to the chelate effect, ligands with two phosphine groups bind more tightly to metal centers than do two monodentate phosphines. The conformational properties of diphosphines makes them especially useful in asymmetric catalysis, e.g. Noyori asymmetric hydrogenation. Several diphosphines have been developed, prominent examples include 1,2-bis(diphenylphosphino)ethane (dppe) and 1,1'-Bis(diphenylphosphino)ferrocene, the trans spanning xantphos and spanphos. The complex dichloro(1,3-bis(diphenylphosphino)propane)nickel is useful in Kumada coupling.

Related Research Articles

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.

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.

<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">Transition metal pincer complex</span>

In chemistry, a transition metal pincer complex is a type of coordination complex with a pincer ligand. Pincer ligands are chelating agents that binds tightly to three adjacent coplanar sites in a meridional configuration. The inflexibility of the pincer-metal interaction confers high thermal stability to the resulting complexes. This stability is in part ascribed to the constrained geometry of the pincer, which inhibits cyclometallation of the organic substituents on the donor sites at each end. In the absence of this effect, cyclometallation is often a significant deactivation process for complexes, in particular limiting their ability to effect C-H bond activation. The organic substituents also define a hydrophobic pocket around the reactive coordination site. Stoichiometric and catalytic applications of pincer complexes have been studied at an accelerating pace since the mid-1970s. Most pincer ligands contain phosphines. Reactions of metal-pincer complexes are localized at three sites perpendicular to the plane of the pincer ligand, although in some cases one arm is hemi-labile and an additional coordination site is generated transiently. Early examples of pincer ligands were anionic with a carbanion as the central donor site and flanking phosphine donors; these compounds are referred to as PCP pincers.

<span class="mw-page-title-main">Triphenylphosphine</span> Chemical compound

Triphenylphosphine (IUPAC name: triphenylphosphane) is a common organophosphorus compound with the formula P(C6H5)3 and often abbreviated to PPh3 or Ph3P. It is versatile compound that is widely used as a reagent in organic synthesis and as a ligand for transition metal complexes, including ones that serve as catalysts in organometallic chemistry. PPh3 exists as relatively air stable, colorless crystals at room temperature. It dissolves in non-polar organic solvents such as benzene and diethyl ether.

<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">Tetrakis(triphenylphosphine)palladium(0)</span> Chemical compound

Tetrakis(triphenylphosphine)palladium(0) (sometimes called quatrotriphenylphosphine palladium) is the chemical compound [Pd(P(C6H5)3)4], often abbreviated Pd(PPh3)4, or rarely PdP4. It is a bright yellow crystalline solid that becomes brown upon decomposition in air.

<span class="mw-page-title-main">Ligand cone angle</span> Measure of the steric bulk of a ligand in a coordination complex

In coordination chemistry, the ligand cone angle (θ) is a measure of the steric bulk of a ligand in a transition metal coordination complex. It is defined as the solid angle formed with the metal at the vertex of a cone and the outermost edge of the van der Waals spheres of the ligand atoms at the perimeter of the base of the cone. Tertiary phosphine ligands are commonly classified using this parameter, but the method can be applied to any ligand. The term cone angle was first introduced by Chadwick A. Tolman, a research chemist at DuPont. Tolman originally developed the method for phosphine ligands in nickel complexes, determining them from measurements of accurate physical models.

Organophosphines are organophosphorus compounds with the formula PRnH3−n, where R is an organic substituent. These compounds can be classified according to the value of n: primary phosphines (n = 1), secondary phosphines (n = 2), tertiary phosphines (n = 3). All adopt pyramidal structures. Organophosphines are generally colorless, lipophilic liquids or solids. The parent of the organophosphines is phosphine (PH3).

<span class="mw-page-title-main">1,1'-Bis(diphenylphosphino)ferrocene</span> Chemical compound

1,1-Bis(diphenylphosphino)ferrocene, commonly abbreviated dppf, is an organophosphorus compound commonly used as a ligand in homogeneous catalysis. It contains a ferrocene moiety in its backbone, and is related to other bridged diphosphines such as 1,2-bis(diphenylphosphino)ethane (dppe).

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.

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

In coordination chemistry, the bite angle is the angle on a central atom between two bonds to a bidentate ligand. This ligand–metal–ligand 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">Organoruthenium chemistry</span>

Organoruthenium chemistry is the chemistry of organometallic compounds containing a carbon to ruthenium chemical bond. Several organoruthenium catalysts are of commercial interest and organoruthenium compounds have been considered for cancer therapy. The chemistry has some stoichiometric similarities with organoiron chemistry, as iron is directly above ruthenium in group 8 of the periodic table. The most important reagents for the introduction of ruthenium are ruthenium(III) chloride and triruthenium dodecacarbonyl.

<span class="mw-page-title-main">Bis(triphenylphosphine)palladium chloride</span> Chemical compound

Bis(triphenylphosphine)palladium chloride is a coordination compound of palladium containing two triphenylphosphine and two chloride ligands. It is a yellow solid that is soluble in some organic solvents. It is used for palladium-catalyzed coupling reactions, e.g. the Sonogashira–Hagihara reaction. The complex is square planar. Many analogous complexes are known with different phosphine ligands.

Organoplatinum chemistry is the chemistry of organometallic compounds containing a carbon to platinum chemical bond, and the study of platinum as a catalyst in organic reactions. Organoplatinum compounds exist in oxidation state 0 to IV, with oxidation state II most abundant. The general order in bond strength is Pt-C (sp) > Pt-O > Pt-N > Pt-C (sp3). Organoplatinum and organopalladium chemistry are similar, but organoplatinum compounds are more stable and therefore less useful as catalysts.

<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.

Phosphinoimidates, also known as phophinimides, are the anions derived from phosphine imides with the structure [R3P=N] (R = alkyl or aryl). Phosphinimide ligands are used to for transition metal complexes that are highly active catalysts in some olefin polymerization reactions.

<span class="mw-page-title-main">Dichlorobis(triphenylphosphine)nickel(II)</span> Chemical compound

Dichlorobis(triphenylphosphine)nickel(II) refers to a pair of metal phosphine complexes with the formula NiCl2[P(C6H5)3]2. The compound exists as two isomers, a paramagnetic dark blue solid and a diamagnetic red solid. These complexes function as catalysts for organic synthesis.

Palladium forms a variety of ionic, coordination, and organopalladium compounds, typically with oxidation state Pd0 or Pd2+. Palladium(III) compounds have also been reported. Palladium compounds are frequently used as catalysts in cross-coupling reactions such as the Sonogashira coupling and Suzuki reaction.

References

  1. Hartwig, J. F. Organotransition Metal Chemistry, from Bonding to Catalysis; University Science Books: New York, 2010. ISBN   1-891389-53-X
  2. Paul C. J. Kamer, Piet W. N. M. van Leeuwen, ed. (2012). Phosphorus(III)Ligands in Homogeneous Catalysis: Design and Synthesis. New York: Wiley. ISBN   978-0-470-66627-2.
  3. Iaroshenko, Viktor (4 January 2019). "Phosphines and Related Tervalent Phosphorus Systems". In Van Leeuwen, Piet W. N. M. (ed.). Organophosphorus Chemistry: From Molecules to Applications. pp. 1–58. doi:10.1002/9783527672240.ch1. ISBN   9783527672240.
  4. Miyaura, Norio; Suzuki, Akira (1993). "Palladium-Catalyzed Reaction of 1-Alkenylboronates with Vinylic Halides: (1Z,3E)-1-Phenyl-1,3-octadiene". Org. Synth. 68: 130. doi:10.15227/orgsyn.068.0130.
  5. 1 2 C. A. McAuliffe, ed. (1973). Transition Metal Complexes of Phosphorus, Arsenic, and Antimony Ligands. J. Wiley. ISBN   0-470-58117-4.
  6. Osborn, J. A.; Wilkinson, G. (1967). "Tris(triphenylphosphine)halorhodium(I)". Inorganic Syntheses. Vol. 10. p. 67. doi:10.1002/9780470132418.ch12. ISBN   9780470131695.
  7. 1 2 Tolman, C. A. (1977). "Steric effects of Phosphorus Ligands in Organometallic Chemistry and Homogeneous Catalysis". Chemical Reviews . 77 (3): 313–348. doi:10.1021/cr60307a002.
  8. Orpen, A. G.; Connelly, N. G. (1990). "Structural Systematics: the Role of P-A σ* Orbitals in Metal-Phosphorus π-Bonding in Redox-Related Pairs of M-PA3 Complexes (A = R, Ar, OR; R = alkyl)". Organometallics . 9 (4): 1206–1210. doi:10.1021/om00118a048.
  9. Gilheany, D. G. (1994). "No d Orbitals but Walsh Diagrams and Maybe Banana Bonds: Chemical Bonding in Phosphines, Phosphine Oxides, and Phosphonium Ylides". Chem. Rev. 94 (5): 1339–1374. doi:10.1021/cr00029a008. PMID   27704785.
  10. Crabtree, Robert H. (2009). The Organometallic Chemistry of the Transition Metals (5th ed.). Wiley. pp. 99–100. ISBN   978-0-470-25762-3.
  11. R. H. Crabtree (2005). "4. Carbonyls, Phosphine Complexes, and Ligand Substitution Reactions". The Organometallic Chemistry of the Transition Metals (4th ed.). Wiley. ISBN   0-471-66256-9.
  12. Newman-Stonebraker, Samuel H.; Smith, Sleight R.; Borowski, Julia E.; Peters, Ellyn; Gensch, Tobias; Johnson, Heather C.; Sigman, Matthew S.; Doyle, Abigail G. (2021). "Univariate classification of phosphine ligation state and reactivity in cross-coupling catalysis". Science. 374 (6565): 301–308. Bibcode:2021Sci...374..301N. doi:10.1126/science.abj4213. PMID   34648340. S2CID   238991361.
  13. Nelson, John H. (2003). Nuclear Magnetic Resonance Spectroscopy. Prentice Hall. ISBN   978-0130334510.
  14. Paul S. Pregosin, Roland W. Kunz (2012). 31P and 13C NMR of Transition Metal Phosphine Complexes. Berlin: Springer. ISBN   9783642488306.
  15. Garrou, Philip E. (1985). "Transition-Metal-Mediated Phosphorus-Carbon Bond Cleavage and Its Relevance to Homogeneous Catalyst Deactivation". Chem. Rev. 85 (3): 171–185. doi:10.1021/cr00067a001.
  16. Reppe, W.; Schweckendiek, W. J. (31 July 1948). "Cyclisierende Polymerisation von Acetylen. III Benzol, Benzolderivate und hydroaromatische Verbindungen". Justus Liebigs Annalen der Chemie. 560 (1): 104–116. doi:10.1002/jlac.19485600104.
  17. Slaugh, L; Mullineaux, R. (1968). "Novel Hydroformylation catalysts". J. Organomet. Chem. 13 (2): 469. doi:10.1016/S0022-328X(00)82775-8.
  18. P. W.N.M. van Leeuwen "Homogeneous Catalysis: Understanding the Art, 2004 Kluwer, Dordrecht. ISBN   1-4020-2000-7
  19. Herrmann, W. A.; Kohlpaintner, C. W. (1998). "Syntheses of Water-Soluble Phosphines and their Transition Metal Complexes". Inorganic Syntheses. Vol. 32. pp. 8–25. doi:10.1002/9780470132630.ch2. ISBN   9780471249214.{{cite book}}: |journal= ignored (help)