Open access peer-reviewed chapter

Versatile Silylphosphine Ligands for Transition Metal Complexation

Written By

Julio Zamora-Moreno and Virginia Montiel-Palma

Submitted: 25 September 2017 Reviewed: 04 January 2018 Published: 23 May 2018

DOI: 10.5772/intechopen.73502

From the Edited Volume

Ligand

Edited by Chandraleka Saravanan and Bhaskar Biswas

Chapter metrics overview

1,702 Chapter Downloads

View Full Metrics

Abstract

In this chapter, a review throughout the literature on the chemistry of multidentate silylphosphines is presented. The incorporation of P and Si functionalities in cooperation in a single ligand backbone is exceptionally versatile, and examples of this rich chemistry stemming from the works of many research groups around the world are herein provided. The ligand systems can be flexible or rigid and incorporate varying numbers of P, Si and even other atoms. Exceptional ligand-metal systems are discussed in terms of their structure, reactivity and, in some cases, catalytic activity.

Keywords

  • silicon
  • phosphorous
  • silylphosphines
  • transition metals
  • multidentate ligands

1. Introduction

In modern Coordination and Organometallic Chemistry, ligand design is recognised as crucial for the development of efficient and selective complexes for important transformations including medicinal chemistry, material science and catalysis. Polydentate-rigid or semi-rigid ligands constrain the geometry at the metal centre providing inherently well-defined coordination geometries for potential incoming substrates. Indeed, a good number of these metal-ligand systems are capable of performing selectively difficult activations and many research groups around the world have directed their endeavours to the study of their chemical properties.

The incorporation of dual functionalities in a single ligand backbone has also been shown to modify the properties of the compounds making them especially prone to undergo selective transformations resulting from differing reactivity of the coordinating atoms in the ligand. A wide variety of combinations of donor atoms have been employed to date, including for example, soft and hard donor atoms in what is known as hemilabile ligands.

In this chapter, the chemistry related to silylphosphine ligands which include in their structure both a basic P as well as a Si is reviewed. Si derivatives are exceptionally good sigma donors and exert a considerably high trans-influence/effect, thus their coordination generates electron rich metal centres in turn capable of activating otherwise inert substrates. Phosphines have long been preferred ligands due to their ability to tune their steric and electronic properties depending on the substituents on P. The incorporation of P and Si in a ligand framework also allows for the employment of NMR spectroscopic tools deriving from 31P and 29Si nuclei.

Advertisement

2. Silylphosphine ligands: definition, general structure and bonding

Silylphosphines can be described as bi- or polydentate ligands bearing at least one basic phosphorous (III) atom, usually a phosphine PR3 or phosphite P(OR)3, and at least one silicon-substituted moiety. The P (III) group is able to form a coordination bond to the transition metal, while the silyl moiety is potentially prone to bind by means of loss of H2, alkane or arene molecules. Between the P and Si atoms, there are generally a number of carbon atoms in the form of an alkyl or aryl bridges (Scheme 1). Else a direct P─Si bond can be established. Silylphosphines are potentially bi-, tri- or polydentate ligands, the coordination number depending on the number of P or Si moieties present in the ligand backbone.

Scheme 1.

General structure of silylphosphine ligands and examples of main coordination modes.

Therefore, the molecular orbitals can be described as those of the phosphine and silicon donor moieties. For example, for the non-classical bidentate coordination mode, the frontier orbitals are shown schematically in Figure 1. A bidentate P, Si ligand can readily coordinate to the metal centre both through the phosphorous atom via the donation of the electron lone pair on P to an empty d-orbital on the metal and through the σ-Si─H electron pair donated to a suitable empty d-orbital on the metal generating a 3c-2e non-classical bond. In both bonds, the stabilisation is given by the retro-donation of electron density of a filled d-orbital to an anti-bonding orbital. In the full oxidative addition process of the Si─H bond to the metal, due to the strong retro-donation of the d-orbital → σ*(Si─H), the final product results in the formation of two 2c-2e bonds: M─H, M─Si. As expected, depending on the substituents on both the P and Si atoms, the molecular orbital diagrams and the energy of the HOMO and LUMO will vary. In general, it could be said according to Figure 1, the HOMO generally possess a higher ligand character, while the LUMO is more metal centred.

Figure 1.

Molecular orbital diagrams of the phosphine and non-classical Si─H sigma moieties.

Advertisement

3. Silylphosphine ligands throughout the chemical literature: a review

Stobart and co-workers pioneered the systematic study of transition metals bound to silylphosphine ligands. As early as in 1983, they reported the synthesis and full characterisation of an extensive family of silanes modified with a phosphorous fragment connected to the silicon atom by a polymethylene chain, of general formula (XYZ)Si(CH2)nPR2 (where X, Y, Z = Me, Ph, Cl or H; n = 1–3; R = Me or Ph;) (Figure 2, compounds 1–33) [1]. The reactivity of Vaska’s complex trans-[Ir(PPh3)2(CO)(Cl)] towards Ph2P(CH2)2SiRR’H (compounds: 12, 14, 16, 18, 20) was also investigated. The reaction results in the formation of air stable six-coordinated Ir(III) compounds, resulting from coordination of the ligands through the P atom and of the oxidative addition of the Si─H bond (compounds: 3438) (Figure 3). Furthermore, the reactions of Ph2P(CH2)2SiMe2H towards the dimers [M(μ-Cl)(COD)]2 (M = Rh, Ir; COD = 1, 5-cyclooctadiene), also afford the M(III) complexes [MCl{Ph2P(CH2)2SiMe2}2] (M = Rh 39; Ir 40) which are quiral with the two phosphorous atoms in trans disposition while the two Si dispose in a cis fashion (Figure 3) [2]. The fact that the reactivity of complexes 39 and 40 was remarkably constrained due to the trans-labilising effect of the silyl groups, was exploited in their use as catalysts for transformations of organic substrates [3]. Several works reported in the literature have argued on the high extent of trans-influence silyl groups exercise on a transition metal centre. There are various reasons for this behaviour including an excellent sigma orbital overlap as well as a favourable electronic release of the Si [4, 5]. This is in agreement with only a few compounds exhibiting a trans coordination of the Si atoms in many cases as kinetic products in equilibria with their cis isomers [6, 7] even when employing chelating silylphosphines (vide supra) [8, 9, 10, 11] (Section 8).

Figure 2.

The silylphosphine ligands bearing alkyl bridges reported in Ref. [1].

Figure 3.

The reactivity of group 9 metals as reported in Ref. [2].

Figure 4.

Design of poly(phosphino)-silane ligands reported in Ref. [12].

It was found that the ligands with two or three phosphorous atoms and a Si-H bond (compounds 41-52, Figure 4) coordinate via oxidative addition to the metal centre (i.e. rhodium, iridium, ruthenium and platinum) and impose steric constraints on the coordination sphere in turn restraining substrate entry to sites which could suffer the strongly labilising trans effect of the silyl group, increasing the complexes’ capabilities as catalysts [12] (Figure 4).

Advertisement

4. Silylphosphines complexation in tetra-coordinated systems

4.1. Square-planar geometry

Turculet and co-workers have further made significant contributions in the field of silylphosphine chemistry. They introduced a PSiP-type ligand [(ο-C6H4-PCy2)2SiMe2] (53) and explored its reactivity with transition metal precursors. The complexes [MCl{(ο-C6H4-PCy2)2SiMe}] (M = Ni or Pd) (64, 65) were treated with alkyl lithium or Grignard reagents. In the case of 65 treatment with stoichiometric amounts of MeLi led to the formation of [Pd(Me){(ο-C6H4-PCy2)2SiMe}] (61) which regenerates complex 65, upon reaction with Ph2SiClH while renders [Pd{(ο-C6H4-PCy2)2SiMe}(SiHPh2)] (66) from reaction with Ph2SiH2. The direct reaction of 53 and [Pd2(dba)3] provides complex [Pd{(ο-C6H4-PCy2)SiMe2}(ο-C6H4-PCy2)] (63) derived from Si-C(sp2) bond activation. Treatment of complex 64 with MeMgBr led to [Ni(Me){(ο-C6H4-PCy2)2SiMe}] (60) and complex [Ni{(ο-C6H4-PCy2)SiMe2}(ο-C6H4-PCy2)] (62) resulting from ligand rearrangement. Complexes 62 (Ni) and 63 (Pd) constitute rare examples of reversible Si-C(sp2) and Si-C(sp3) bond cleavage (Figure 5) [13, 14].

Figure 5.

[PSiP] ligands and their square planar group 10 metal complexes [13-19, 26, 59].

Moreover, the reactivity of 54 towards [PtCl2(SEt2)2] leads to the generation of [PtCl{(ο-C6H4-PPh2)2SiMe}] (67) where the ligand coordination results in adoption of a distorted square planar geometry around Pt with a persistent Cl atom bonded trans to the silyl group (Figure 5) [9].

Iwasawa and collaborators reported an interesting system for the catalytic hydrocarboxylation of allenes using the Pd(II) hydride complex [PdH{(ο-C6H4-PPh2)2SiMe}] (68) as the active catalyst. Their methodology also served for the synthesis of β,γ-unsaturated carboxylic acids. In general, complex [Pd(OTf){(ο-C6H4-PPh2)2SiMe}] (69) in catalytic ratios 1.0 mol% or 2.5 mol % was used in soft conditions of CO2 pressure (1 atm) with 150 mol% of AlEt3 or ZnEt2 for carboxylation of 1,1-disubstituted, monosubstituted or disubstituted allenes to the respective carboxylic acid or ester [15]. In addition, the chemical properties of complex 69 were described as well. To mention some, 69 undergoes transmetalation with AlEt3 followed by β-hydride elimination to generate the proposed complex 68. Complex 69 reacted with an excess of B2pin2 at room temperature leading to HBpin and the monoborylpalladium complex [Pd(Bpin){(ο-C6H4-PPh2)2SiMe}] (70) which promotes the product of borylation of styrene as well as other alkenes (Figure 5) [16, 17, 18].

Milstein and co-workers described the design and synthesis of the first pincer-type silanol-Pt(II) compound by using a PSiP ligand. The ligand {(ο-C6H4-PiPr2)2SiH2} (58) was obtained in moderate yields from the o-bromophosphine. It readily reacts with [(Me2S)2Pt(Me)Cl] at room temperature to give the bicyclic complex [PtCl{(ο-C6H4-PiPr2)2SiH}] (72), which then undergoes autoxidation yielding the silanol complex [PtCl{(ο-C6H4-PiPr2)2Si(OH)}] (73) in moderate yields (65%) (Figure 5) [19].

Interestingly, changes on the identity of the substituents on the P atoms in the PSiP ligand backbone bring about a great strategy for the coordination of ligand 55 towards group 9 metals, in particularly rendering an Ir system able to activate intermolecular arene C─H bonds. Indeed, the reaction of 55 and [MCl(coe)2]2 (M = Rh, Ir; coe = 1-cyclooctadiene) or [RhCl(PPh3)3] produced the monomeric complexes [MH(Cl){(ο-C6H4-PCy2)2SiMe}] (M = Rh, 74; Ir, 75), which react with organolithium compound [Me3SiCH2Li] forming neutral three-coordinate intermediate species able to subsequently coordinate neutral ligands, thus generating [M(L){(ο-C6H4-PCy2)2SiMe}] (M = Rh, L = H2NPh (76); M = Rh, L = NH3 (77); M = Ir, L = C2H4 (78); M = Ir, L = PMe3 (79); M = Rh, L = PMe3 (80)) (Figure 7) [20].

4.2. Tetrahedral and trigonal pyramidal geometries

Ligand 54 (see Figure 5) reacted with [Pd(PPh3)4] or [CpPd(C3H5)] yielding complex [Pd{η2-(ο-C6H4-PPh2)2SiHMe}(PPh3)] (82) instead of the hypothesised hydride complex 68 which was proposed as the intermediate in the reduction of compound [PdCl{(ο-C6H4-PPh2)2SiMe}] (71, see Figure 5) with LiHBEt3 in presence of PPh3 to afford also 82 [21]. Likewise, the reaction of ligand 54 with [Ni(PPh3)4] led to the formation of the Ni(0) complex [Ni{η2-(ο-C6H4-PPh2)2SiHMe}(PPh3)] (81). The Si─H, Ni─H and Ni─Si distances are 1.62(3), 1.44(2) and 2.2782(4) Å respectively, suggesting that the Si─H bond was preserved. This non-classical complexation mode is kept in solution because the observed NMR parameters such as coupling constants JSiH = 89 Hz at 300 K and 77 Hz at 193 K are large (2JSiH ≤ 20 Hz for a complete Si─H bond cleavage) and in line with the conservation of the η2-Ni(0) structure seen in solid state (Figure 6). On the other hand, the mixture of 54 with [Pt(PPh3)4] is discussed in Section 5.2 [22].

Figure 6.

Reactivity of 54. Tetra-coordination is highlighted in blue and penta-coordinated in purple [21-23, 26-29].

Bourissou and co-workers reported the reactivity of the PSiP ligand 54 towards CuCl and AuCl(SMe2), which was subsequently followed by a stoichiometric addition of GaCl3 (complexes 84, 85). The addition of the gallium halide was envisioned to increase the electrophilicity of the central metal and thus to escalate the strength of non-classical σ-SiH bond interaction at the metal. In complex 84, the coordination of the ligand occurs through the two phosphorous atoms and a weak sigma interaction Si─H⋯Cu. The spectroscopic evidence as well as computational analyses (geometry optimisations and NBO analyses) are in agreement with weak donation σ-SiH → Cu in combination with a negligible Cu → σ*SiH back-donation in 84. Meanwhile in the cationic gold complex 85, the coordination of 54 took place only through the two phosphorous atoms as any non-classical Si─H bond interaction to the metal was strongly disfavoured as it was found to be by computational means 15.9 kcal/mol (Figure 6) [23].

Extraordinarily, 86 (described in more detail in Section 5.2) demonstrated to be a suitable precursor for the synthesis of stable 14-electron [Ru(X){(ο-C6H4-Cy2)2SiMe}] (X = OtBu (87), N(SiMe3)2 (88); NHPh (89); NH(2,6-Me2C6H3) (90)) complexes, donning unusual trigonal pyramidal coordination geometries explained once again by the presence of the strongly σ-donating silyl group in the apical site with the contribution of steric effects of the phosphino substituents in the equatorial plane (Figure 7) [24].

Figure 7.

Reactivity of ligand 55. Tetra-coordination is highlighted in blue and penta-coordination in purple [20, 24-25, 36].

Advertisement

5. Silylphosphines complexation in penta-coordinated systems

5.1. Square pyramidal geometry

The reactivity of 55 with [RuCl2(p-cymene)]2 was carried out in the presence of PCy3 and Et3N rendering a binuclear complex that preserves bridging chloride ligands [Ru(μ-Cl){(ο-C6H4-PCy2)2SiMe}]2 (86), which was exhaustively characterised spectroscopically (Figure 7). The ligand coordinates each Ru atom through two P, one Si and two Cl atoms in a distorted square pyramidal geometry, in which the silyl group occupies the apical coordination site. This Ru(II)-ligand system brings about stability and selectivity in catalytic transformations including the transfer hydrogenation of ketones. In this case, Li et al observed high conversion of the corresponding secondary alcohols for numerous dialkyl, diaryl, and alkyl/aryl ketones, employing 0.2 mol% of 86 with 5 mol% of KOtBu at 80°C [25].

The activation of Si-H bonds in ligands of general formula (ο-C6H4-PR2)2SiHMe (R = Ph or Cy; 54 and 55) was also observed in monomeric Ru systems. Compound 54 reacted with [RuCl2(PPh3)3] in the presence of triethylamine as a base affording complex [RuCl{(ο-C6H4-PPh2)2SiMe}(PPh3)] (91). The X-ray diffraction study confirms the coordination of the ligand through the two phosphorous and the silicon atom adopting a facial arrangement with the silyl group occupying the basal position of a distorted square pyramidal geometry around Ru. The Cl atom disposes trans to the silyl group, and the remaining site was occupied by a PPh3. Complex 91 reacted with LiEt3BH to form octahedral-Ru hydride complex 92 in moderate yield (Figures 6) [26].

5.2. Trigonal bipyramidal geometry

The mixture of 54 with [Pt(PPh3)4] at room temperature led to the generation of five-coordinated Pt(II) complex [PtH{(ο-C6H4-PPh2)2SiMe}(PPh3)] (93). In contrast with the derivatives of Ni(0) 81 and Pd(0) 82 (Figure 6), where the Si─H bond is only slightly activated, the Pt(II)-hydride complex 93 derives from the complete oxidative addition of the Si─H bond. The crystalline structure displays a trigonal bipyramidal geometry with the silyl group in the apical position in the metal centre. The opposed apical site was taken by the hydride ligand which in the 1H NMR spectrum revealed a quartet at δ −7.92 ppm (2JPH = 18.9 Hz) with 195Pt satellites exhibiting a measured coupling constant 195Pt-1H of 650 Hz, which is considerably small compared with some cis-H-Pt(II)-Si species previously reported (1JPtH = 890–1010 Hz); supporting the proposal that the hydride is located trans to Si atom (Figure 6) [22].

A study of the reaction of complex 82 and its related analogue [Pd{η2-(ο-C6H4-PPh2)2SiHMe}(PMe3)] (83) towards B2pin2 was made, since it could provide a means of accessing Pd(II) hydrides via oxidative addition of the Si─H bond. Two isomers: cis and trans were proposed. Depending on the phosphine choice, the isolation of one isomer was possible through a reversible σ-bond metathesis pathway. In the case of the PMe3 ligated complex, the kinetic product cis-[Pd(Bpin){(ο-C6H4-PPh2)2SiMe}(PMe3)] (94) showed a slow reverse reaction and was obtained predominantly. In contrast, for the PPh3 derivative, the equilibrium favoured the thermodynamic isomer trans-[Pd(Bpin){(ο-C6H4-PPh2)2SiMe}(PPh3)] (95) as a major product (Figure 6) [27, 28]. In relation to unusual bonding modes, Sun and collaborators reported the systematic reactivity of the tridentate ligand 54 towards the low-valent nickel compound [Ni(PMe3)4] which induced the formation of Ni(0) complexes [Ni{η2-(ο-C6H4-PPh2)2MeSi-H}(PMe3)] (96, Figure 6). Complex 96 did not undergo the oxidative addition process of a Si─H bond even in the presence of independent silanes (Et3SiH, Ph2MeSiH). However, the reactivity with chlorosilanes Me3SiCl or MeHSiCl2 led to the formation of Ni(II) complex [NiCl{(ο-C6H4-PPh2)2MeSi}] (97). The halogenated products [NiX{(ο-C6H4-PPh2)2SiMe}(PMe3)] (X = Br (98); I (99)) were easily obtained from reaction with EtBr or MeI of complex 96 (Figure 6) [29].

The versatility of ligand 55 was also probed in the coordination towards Ru. With the aim of preparing highly valuable 16-electron complexes, complex 86 was reacted with monodentate phosphines. The reaction with PPh3 results in small conversion to the five-coordinated compound [Ru(Cl){(ο-C6H4-PCy2)2SiMe}(PPh3)] (100) in equilibrium with 86. Interestingly, this latter compounds are also formed from the reaction of ligand 55 and [RuCl2(PPh3)3] in the presence of NEt3 albeit in low yields. However, the production of the isolable penta-coordinate complex [Ru(Cl){(ο-C6H4-PCy2)2SiMe}(PMe3)] (101) was possible in quantitative yields when employing 86 in solution and the smaller, more σ-electron-donating PMe3 (Figure 7) [30].

In complexes 74 and 75 (Figure 7), the ligand coordinates in a tridentate fashion through the phosphorous atoms which dispose in trans and the Si which sits in the equatorial plane of a trigonal bipyramidal geometry. The remaining sites were taken by the hydride derived from the ligand and a Cl atom. Remarkably, besides the intermolecular C─H activation ability imposed by the coordination of ligand 55 to Ir, complex 75 also exhibits facile N─H bond activation of ammonia and anilines while its Rh analogues undergo mainly adduct formation. Likewise, the starting complexes 74 and 75 react with lithium anilides [Li(NHR)] generating isolable anilido hydride complexes [MH(NHR){(ο-C6H4-PCy2)2SiMe}] (M = Rh, R = Ph (102); M = Rh, R = {2, 6-Me2C6H3} (103); M = Ir, R = Ph (104); M = Ir, R = {2, 6-Me2C6H3} (105); M = Ir, R = H (106)) upon mixing. The new compounds were described as being very resistant to N─H bond reductive elimination even in the presence of alkyl or aryl substrates (Figure 7) [20].

The ligand (ο-C6H4-PPh2)3Si-H (107) reacts with [Ni(PPh3)4] to yield the complex [Ni{η2-(ο-C6H4-PPh2)2Si-H(ο-C6H4-PPh2)}(PMe3)] (109) bearing non-classical σ-Si─H bonds. On the other hand, complex 109 undergoes thermal oxidative addition at the Ni centre and loss of PMe3 to allow the coordination of the previously uncoordinated phosphorous, thus rendering a compound of formula [NiH{(ο-C6H4-PPh2)3Si)}] (110). In a subsequent step, HCl was added to afford the formation of [NiCl{(ο-C6H4-PPh2)3Si)}] (111), which was also obtained when compound 109 was combined with one equivalent of MeHSiCl2. Compounds [NiX{(ο-C6H4-PPh2)3Si}] (X = Br, 112; I, 113) were obtained from the reaction of 110 with either EtBr or MeI (Figure 8).

Figure 8.

The chemistry of P3Si systems [31-34].

Peters and co-workers have also reported the synthesis and reactivity of silanes functionalised with phosphines and/or sulphur derivatives. In particular, the ligand 107 reacts with [Fe2Mes4] leading to the formation of [Fe{(ο-C6H4-PPh2)3Si}Mes] (114), which was characterised structurally by single crystal X-ray diffraction. The analysis discloses a distorted octahedral geometry around the Fe atom in which the ligand has taken four out of the six coordination positions, a mesityl group occupies one more and the sixth site (trans to silyl group) is occupied by an agostic interaction (C─H⋯Fe) from a methyl group in ortho position of the mesityl bonded to the metal. The reaction of this agostic complex with HCl leads to the formation of complex [FeCl{(ο-C6H4-PPh2)3Si}] (115). Subsequent reduction with Na/Hg amalgam under a N2 atmosphere led to the Fe(I) complex [Fe(N2){(ο-C6H4-PPh2)3Si}] (116), which was again subjected to reduction with an additional equivalent of Na/Hg in the presence of [12]crown-4 to render dark purple ionic pair [Na([12]crown-4)2]+[Fe(N2){(ο-C6H4-PPh2)3Si}] (117) in which Fe is in a zero oxidation state and the dinitrogen ligand is less labile than in 116 because of stronger π backdonation from the more reduced metal [32]. Once again the nature of the substituents on the P atom is determinant. Indeed, ligand (ο-C6H4-PiPr2)3SiH (108) reacts with FeCl2 at room temperature producing a species with one uncoordinated phosphorous atom and the Si─H bond intact [FeCl2{(ο-C6H4-PiPr2)2SiH(ο-C6H4-PiPr2)}] (118); however, if the reaction was made in the presence of MeMgCl at −78°C, the desired complex [FeCl{(ο-C6H4-PiPr2)3Si}] (119) was isolated. The reactivity of 108 with metallic precursors CoCl2, NiCl2 and [Ir(μ-Cl)(COD)]2 was found to yield the tripodal species [CoCl{(ο-C6H4-PiPr2)3Si}] (120), [NiCl{(ο-C6H4-PiPr2)3Si}] (121) and [IrHCl{(ο-C6H4-PiPr2)3Si}] (122), respectively (Figure 8) [33].

Another example of the importance of the trans-influence of the silyl groups on the metal coordination sphere is given by the chemical properties of [RuCl{(ο-C6H4-PPh2)3Si}] (123). The exchange reaction with LiPPh2 (also with LiPiPr2) led to expected phosphide complex [Ru(PPh2){(ο-C6H4-PPh2)3Si}] (124), which decays at room temperature to the cyclometalated [Ru(PHPh2){Si(ο-C6H4-PPh2)2(ο-C6H4)P(ο-C6H4)Ph}] (125). The reaction of 123 with stoichiometric amounts of MeLi or PhCH2MgCl leads to [Ru(Z){(ο-C6H4-PPh2)3Si}] (Z = Me (126); PhCH2 (127)). Successive loss of methane or toluene renders the formation of the unsaturated square pyramidal species [Ru{Si(ο-C6H4-PPh2)2(ο-C6H4)P(ο-C6H4)Ph}] (128), which in turns affords under H2 or N2 atmospheres compounds [Ru(H2)H{(ο-C6H4-PPh2)3Si}] (129) and [Ru(N2)H{(ο-C6H4-PPh2)3Si}] (130) in subsequent reactions steps (Figure 8) [34].

Advertisement

6. Silylphosphines complexation in hexa-coordinated systems: octahedral geometry

Shimada and collaborators reported on the reactivity of ligands (ο-C6H4-PR2)2SiHMe (R = Cy (55), iPr (56), tBu (57); Figure 5) towards [Ir(μ-Cl)(COD)]2. The complexes [IrClH{(ο-C6H4-PR2)2SiMe}] (R = Cy, 75, iPr, 131, tBu, 132) (Figure 7 and 9) reacted with reducing agent Me4N·BH4 under argon to produce the tetrahydride complexes [IrH4{(ο-C6H4-PR2)2SiMe}] (R = Cy (133), iPr (134), tBu(135)) or under dinitrogen gas to produce rare stable Ir(III) dihydride-dinitrogen complexes of formula [IrH2(N2){(ο-C6H4-PR2)2SiMe}] (R = Cy (136), iPr (137)). For the last two complexes, NMR spectroscopy reveals the presence of the fac/mer isomers; the meridional and facial disposition of the PSiP ligand was supported by single crystal X-ray diffraction (Figure 9) [35].

Figure 9.

(Left) Chemistry of PSiP complexes of group 9. (Right) Related Fe octahedral complexes [35].

The complex [FeH{(ο-C6H4-PPh2)2SiMe}(PMe3)2] (138) was synthesised from [Fe(PMe3)4], its ν(Fe─H) stretching band was found at 1870 cm−1 in the IR spectrum, while a triplet of doublets (td),signal was found at δ −17.09 (JPH = 71.4, 18.1 Hz) in the 1H NMR spectrum. By comparison, complex [FeH{(ο-C6H4-PPh2)3Si}(PMe3)] (139) was characterised by a ν(Fe─H) band at 1967 cm−1 and a signal in 1H NMR spectrum at δ −15.00 of a triplet of doublets of doublets (tdd) multiplicity with JPH = 79.4, 78.8 and 10.8 Hz (Figure 9) [31].

Advertisement

7. Hybrid silylphosphines complexation

PSiN-ligated complexes have also been attractive synthetic targets due to the hemilability property the presence of soft P, Si and hard N atoms could impose. The synthetic strategy for the preparation of PSiN-type ligands involved the synthesis of {(ο-C6H4)-NMe2}SiHMeCl and its reactivity towards organolithium {(ο-C6H4Li)-PtBu2}. In this manner, the mixed-donor ligand {(ο-C6H4)-PtBu2}{(ο-C6H4)-NMe2}SiHMe (140) was synthesised. This new-fangled ligand reacted with group 10 precursors PdBr2 or [PtCl2(cod)] to give the complexes [M(X){(ο-C6H4)-PtBu2}{(ο-C6H4)-NMe2-SiMe}] (M = Pd, X = Br (144); M = Pt, X = Cl (145)), respectively. Treatment of the latter complexes with OTf- led to the formation of compounds [M(OTf){(ο-C6H4)-PtBu2}{(ο-C6H4)-NMe2-SiMe}] (M = Pd (146); Pt (147)) showing in the X-ray diffraction molecular structure a distance Pd-O of 2.3518(11) Å in 146, which once again highlights the strong trans-influence of the silyl donor. The selective reversible de-coordination of the amine arm in 144 and 145 was observed upon the use of PMe3 which yields compounds [M(X){(ο-C6H4)-PtBu2-SiMe-(ο-C6H4)-NMe2}(PMe3)] (M = Pd, X = Br (148); M = Pt, X = Cl (149)). The coordination of the PSiN ligand towards group 9 (Rh, Ir) and 8 (Ru) has also been studied. The complex [Ru{(ο-C6H4)-PtBu2}{(ο-C6H4)-CHNMe-SiMe}(η3-cyclooctene)] (150) was achieved upon thermal reaction of the PSiN ligand with one equivalent of [Ru(2-methylallyl)2(cod)]. The complex resulted from the coordination of the P and Si atoms of the ligand as well as a C─H bond activation of the methyl group (NMe) with a hydrogenated cyclooctene, remaining on the coordination sphere of the metal. Overall, a square planar geometry around Ru centre is structurally proposed (Figure 10) [36].

Figure 10.

The chemistry of mixed PSiN ligands [36-37].

Another example of an elegant catalytic application of systems derived of PSiN pincer-like ligands is that comprising the ligands of general formula {(ο-C6H4)-PR2}{(ο-C6H4)-NMe2}SiHMe (R = tBu (140); Cy (141); iPr (142); Ph (143)) [36]. The complexes [PtCl{(ο-C6H4)-PR2(ο-C6H4)-NMe2-SiMe}] (R = tBu (145); Cy (151); iPr (152); Ph (153)) were synthesised by the reaction with [PtCl2η2-(C2H4)]2 in the presence of Et3N. In particular, the PSiN-platinum complex 151 successfully catalysed C─H borylation not only of highly electron deficient perfluoroarenes but also of the monofluorinated arenes, chloroarenes and benzoate (Figure 10) [37].

Regarding [M(P3Si)] tripodal systems, an extended series of organometallic species of general formulae [M(X){(ο-C6H4-PR2)3Si}] or [M(L){(ο-C6H4-PR2)3Si}] (R = Ph, iPr; X = Me, Cl; L = H2, N2, CO, NH3, N3R’, PMe3) has been disclosed with a variety of metals including, Fe [38, 39, 40, 41, 42, 43, 44], Co [45, 46], Ni [47], Ru [48], Os [48], Rh [45] and Ir [45], with outstanding and specific properties. Perhaps among the most novel systems, one can find the chemistry of novel hybrids (thioether/phosphine)-silane ligands (ο-C6H4-PiPr2)3-n(ο-C6H4-SAd)nSiH (n = 2 (154); 1 (155)) synthesised from silyl-phosphines such as (ο-C6H4-PiPr2)2SiHCl, 59 (Figure 5, section 4.1), upon lithiation of Br(ο-C6H4)E (E = PiPr2 or SAd) with varying stoichiometric amounts of Li(ο-C6H4)E. The reactivity of these hybrids with FeCl2 afforded a new class of iron complexes featuring a S─Fe─N2 linkage (Figure 11) [49]. The bulky hexa-dentate ligand {(ο-C6H4-PiPr2)2HSi-O-SiH(ο-C6H4-PiPr2)2} (156) was synthesised by the controlled hydrolysis of 59 (Figure 11). The reactivity of 59 was also reported towards FeBr2 and NiX2·DME (X = Cl or Br). The complex [Fe2(N2)(μ-H)2{((ο-C6H4-PiPr2)2Si)2O}] (157) was formed in an equilibrium mixture with [Fe2(N2)2(μ-H)2{((ο-C6H4-PiPr2)2Si)2O}] (158), which were observed at low temperature in the IR spectrum (two ν(N─N) bands at 2097 and 2060 cm−1) in accordance with the determined thermodynamic parameters including a large negative entropy (−30(2) cal/mol K), consistent with the coordination of a gas molecule and a rather small enthalpy of binding (−9.0(4) kcal/mol) in line with the observation of both species at low temperature (Figure 11) [50]. This research is particularly relevant for the understanding of nitrogenase mimicking systems. From the reactivity of a binucleating variant of a PSiP ligand with NiX2.DME (X = Cl, Br) in the presence of triethylamine, dinuclear zerovalent nickel complexes bearing both η2-(Si─H) and η2-H2 moieties were observed by the group of Peters. Theoretical studies suggest that the Ni centre facilitates H atom exchange between the η2-(Si─H) and η2-H2 ligands via interconversion with a higher valent Ni(II) isomer (compounds 159161) (Figure 11). This exchange has been exploited in the selective catalytic deuteration of exogenous silanes [51].

Figure 11.

Chemistry of compound 59 and related reactions [49-52].

Interestingly, the synthesis of the bulkier ligand (ο-C6H4-PiPr2)2HSi-(C6H3)SiPr-SiH(ο-C6H4-PiPr2)2 (162) was reported very recently. The lithiation of (2,6-dibromophenyl)-isopropyl thioether with nBuLi in one pot followed by the stoichiometric addition of 59 affords the formation of (3-bromo-2-(isopropylthio)phenyl)(bis(2-diisopropylphenylphosphino)silane, which received the same treatment that (2,6-dibromophenyl)-isopropyl, to form the desired compound 162 (Figure 11) [52].

Advertisement

8. Bulky silylphosphines complexation

An example of rare kinetic stabilisation of trans bis(silyl) isomers was provided by the contributions of Kang, Ko and coworkers on the reaction of the bulky carborane silyl-phosphines {(R2P)C2B10H10(SiMe2H)} (R = Me (163), OEt (164), Ph (165)) towards [Pt(η2-C2H4)(PPh3)3] or [Pt(COD)2], which afforded extremely uncommon trans-bis(P,Si-chelates) [Pt{(R2P)C2B10H10(SiMe2)}2] (R = Me (166); OEt (167)) formed by “chelate-assisted” oxidative addition. However, in the presence of dimethyl acetylenedicarboxylate, the complexes endure thermally rearrangements to the thermodynamically favoured cis isomers 166′ and 167′. Besides, the reaction of [Pt(η2-C2H4)(PPh3)3] towards 165 occurs via oxidative addition resulting in the mono(chelate) [PtH{(Ph2P)C2B10H10(SiMe2)}(PPh3)] (168). These authors also investigated the reactivity of the same bulky carborane silyl-phosphines with the palladium precursor [Pd2(dba)3] and observed that depending of the substituents over the phosphorous atoms, the cis isomer was exclusively formed [Pd{(R2P)C2B10H10(SiMe2)}2] (R = OEt (169); Ph (170)) or a mixture of the trans/cis isomers was detected when R = Me (171/171′) (Figure 12) [53].

Figure 12.

Bulky carborane silyl-phosphine ligands [53].

Recently, the syntheses of bulky-cage trigonal bipyramidal iron complexes 174 and 175 with remote tertiary amines were reported. The synthesis of ligands 172 and 173 is shown in Figure 13. Once again, in this regard, the incorporation of secondary sphere interactions into iron-phosphine scaffolds is relevant to synthetic nitrogen fixation research [54].

Figure 13.

Syntheses of bulky cage trigonal bipyramidal iron complexes [54].

Advertisement

9. Non-rigid and semi-rigid silylphosphines

Sola reported tridentate systems exemplified by [IrHCl{[Ph2P(CH2)3]2SiMe}] (176) [55] derived from the reaction of the ligand PSiP {[Ph2P(CH2)3]2SiHMe} (47, Figure 3) with the dimeric compound [Ir(μ-Cl)(cod)]2. In solution, complex 176 displays an equilibrium between the syn (176) and anti (176′) isomers in a ratio 93:7 in C6D6 and C7D8, while in CDCl3 or CD2Cl2 solutions, the ratio is ca. 83:17. Complex 176 (and 176′) reacted with NaX (X = Br or I) leading to the corresponding complexes [IrHX{[Ph2P(CH2)3]2SiMe}] ((X = Br (177); I (178)) also in equilibrium with their respective syn and anti-isomers (177′, 178′) in similar ratios that those of 176 [56]. The mixture of isomers 176, also reacted with Me(O3SCF3) to produce the isomers syn179 and anti179′ with general formula [IrH(O3SCF3){[Ph2P(CH2)3]2SiMe}]; likewise the reactivity of 176 with AgX or HX (X = PF6) in the presence of a neutral ligand afforded the mixture of the respective syn/anti cationic species [IrH(L)2{[Ph2P(CH2)3]2SiMe}]+[PF6] (L = NCMe (180/180′), CO (181/181′), bipy (182/182′)) (Figure 14) [57].

Figure 14.

The chemistry of [IrP2Si] [56, 57].

Our research group studied the reactivity of PSi ligand phosphino-(benzyl)-silane Ph2P{(ο-C6H4)CH2SiMe2H} (183) towards the complexes [RuH2(η2-H2)(PCy3)2] (184) and [Ru(cod)(cot)]. Complex 185 resulted from the substitution of two molecules of dihydrogen and two of the ligands PCy3 in 184 by two ligands 183 bonded to the ruthenium atom through the phosphorous atoms and two σ-bonds of the fragments Si─H. Following loss of H2, complex 183 slowly transformed to the cyclometalated complex 186 and subsequently into the bis(cyclometalated) 187. When 183 was added to [Ru(cod)(cot)], the synthetic precursor of 184, it generated directly complex 187 in very high yield. The increase on the acidity of the methylene groups of ligand 183 with respect to a non-benzilic phenylphosphine analogue, coupled with the presence of the non-classical Si─H bond interactions, which could undergo a low energy dissociation-coordination process of the Si─H bonds, was claimed to induce the gradual loss of H2 in 185 to the final stable bis(carbometallated) complex 187. Thus, it was reasonable to propose that the agostic interactions preceded and favoured the C─H bonds activation process [58]. Ligand SiPSi phosphinodibenzyl-silane PhP{(ο-C6H4)CH2SiMe2H}2 (188) was synthesised from PhP(ο-tolyl)2, it behaved as a pincer-like ligand capable of adopting different coordination modes at ruthenium through different degrees of Si─H bond activation. The reaction of 188 towards complex 184 yielded exclusively the formation of 189, in which a Ru(II) centre is coordinated to one ligand 188, through the P atom and two non-fully activated Si─H bonds preserving one PCy3 and two hydride ligands of the original Ru complex. The phosphorous atoms arrange in a distorted cis with a P-Ru-P angle 113.32(4)° in 189 which should be compared to 107.1(4)° in bis-cyclometallated 187. This sterically encumbered arrangement of the phosphine ligands around ruthenium has been explained due to the favourable exchange of the two formally terminal and two non-classical sigma hydrides around the metal. Certainly, the measured value of the JSiH together with theoretical calculations and the observed chemical behaviour of 189 in solution agree with the presence of non-classical η2-Si-H character of the silyl moieties. Thus, the complex 189 was formulated as an 18-electron species stabilised by two unusual intramolecular ε-non-classical interactions. Complex 189 undergoes facile and reversible loss of dihydrogen to afford quantitatively 16-electron complex 190, which is thought to preserve a single non-classical hydride as well as a terminal one. Moreover, NMR spectroscopic experiments on complex 189 show it to be very fluxional in the temperature range accessible, while hydride exchange in complex 190 takes place at the high-temperature regime but in the slow exchange indicates only one hydride is bound to the two silicon atoms. In the solid-state X-ray diffraction analysis, a P-Ru-P angle of 154.37(3)° was determined, which is significantly more obtuse than in 189 and in 187, as expected due to the diminished hydride exchange in 190. The free energy ΔrG298 of the reaction 189 to 190 + H2 is +16.9 kJ/mol; in line with the experimentally observed conversion at 308 K [59]. The reactivity of the ligands 183, 188 and new PSi3 ligand P{(ο-C6H4)CH2SiMe2H}3 (191) was also investigated with compounds [M(μ-Cl)(cod)]2 (M = Rh, Ir) [60] and with [Pt(PPh3)3] [8]. Compound 191 coordinates to Rh and Ir centres as a tetradentate ligand through the phosphorus and two silyl groups, while a third Si atom engages in an agostic Si─H interaction mode [60]. Complexes 192 and 193 react with adventitious water to generate dimeric siloxane compounds. Additionally, compounds 188 and 191 react with Pt as tridentate ligands leading exclusively to compounds exhibiting a very rare trans silyl disposition at square planar Pt (194, 195). These two complexes feature ligand (188 and 191) in a close to meridional disposition. Complex 195 results from ligand modification at one of the benzylic positions which undergoes formation of a new C-Si bond. Furthermore, d8 Pt(II) complex 195 is the first case of a silyl-platinum complex that includes a novel C─H⋯Pt anagostic interaction (Figure 15).

Figure 15.

Chemistry of silyl-benzyl phosphines bi-, tri- and tetradentate [58-60].

Advertisement

10. Applications of silylphosphines in the chemical industry

From the examples throughout this chapter, one can safely envisage transition metal complexes of silylphosphines as active catalysts in a variety of industrial processes. The industrial application of this type of ligand systems, nevertheless, is still at its cradle with future applications expected to materialise in the mid-term.

In principle, Si and P are capable of displaying nucleophilic behaviour and both also possess the ability to displace leaving groups such as halogens, neutral/monodentate ligands, and so on, while the factors affecting their stereochemistry may also assist the complex in the attainment of specific geometries [61]. Catalysed transfer hydrogenation has been developed mainly based on complexes derived from the platinum-metals group [62], and it is applied in industrial process and organic synthesis [63]. [PSiP-Ru] species also have shown to play an excellent role in the reduction of ketones employing iPrOH as the hydrogen source. The well-known Kumada’s cross-coupling reaction is an actual tool for the low-cost synthesis of styrene derivatives in the industrial scale by using Ni and Pd complexes as catalysts [64]. Some advances revealed the crucial use of phosphorous-containing compounds [65, 66, 67] and/or the very bulky donor ligands [68, 69]. Nevertheless, [PSiP-Co] systems have shown efficient conversions in relative mild reaction conditions of an aryl-Grignard reagent reaction with organic halides at 50°C for 24 h [70].

11. Conclusion and perspectives

The incorporation of dual functionalities P and Si in single ligand backbones, silylphosphines, notably modifies the properties of the complexes they form, making them especially reactive and able to undergo selective transformations resulting from differing reactivity of the coordinating atoms in the ligand in conjunction with the chelate effect.

Predictably, the observed reactivity stems from the combination of the most important qualities of the Si ligands, specifically their extremely high σ-donating character and thus their capability of forming σ-complexes, coupled to those features of the P moieties, which can be greatly modified by the choice of substituents.

Throughout this chapter, it has been shown the study of transition metal systems bonded to silylphosphine ligands has thrived in the last decades, but the findings in the last years highlight the importance of their study. Numerous extraordinary systems displaying unusual bonding modes, structures or physicochemical properties have been reported to date and many more can be envisioned to be informed in the near future given the relatively accessible synthesis of ligands and the seemingly unlimited structural variations.

However, the catalytic and other applications of these compounds have been sparingly explored; yet the potential of many of the reported systems is foreseen. We thus expect this field of chemistry to continue growing rapidly and encourage other research groups to direct their endeavours to this fascinating area of research.

Acknowledgments

We acknowledge financial support from CONACyT-Mexico through project CB 242818 and ANR-CONACyT 274001 and a PhD grant to JZ-M

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Holmes-Smith RD, Osei RD, Stobart SR. Phosphinoalkylsilanes: Synthesis and spectroscopic properties of phosphino(silyl)methanes, 1-phosphino-2-silylethanes, and 1-phosphino-3-silylpropanes. Journal of the Chemical Society, Perkin Transactions. 1983; I:861-866
  2. 2. Auburn MJ, Holmes-Smith RD, Stobart SR. (Phosphinoalkyl)silyl complexes: 3. “Chelate-assisted” hydrosilylation: Formation of enantiomeric and diasteroisomeric iridium(III) complexes with chelating (phosphinoethyl)silyl ligands. Journal of the American Chemical Society. 1984;106:1314-1318
  3. 3. Auburn MJ, Stobart SR. (Phosphinoalkyl)silyl complexes. 5. Synthesis and reactivity of congeneric chelate-stabilized disilyl complexes of rhodium(III) and iridium (III): Cholorobis[[(diphenylphosphino)ethyl]dimethylsilyl] rhodium and -iridium. Inorganic Chemistry. 1985;24(3):318-323
  4. 4. Zhu J, Lin Z, Marder TB. Trans influence of boryl ligands and comparison with C, Si, and Sn ligands. Inorganic Chemistry. 2005;44(25):9384-9390
  5. 5. Koller SG, Martín-Romo R, Melero JS, Colquhoun VP, Schildbach D, Strohmann C, et al. Structural consequences of an extreme difference between the trans influence of the donor atoms in a palladacycle. Organometallics. 2014;33(24):7329-7332
  6. 6. Kim Y-J, Park J-I, Lee S-C, Osakada K, Tanabe M, Choi J-C, et al. Cis and trans isomers of Pt(SiHAr2)2(PR3)2 (R = Me, Et) in the solid state and in solutions. Organometallics. 1999; 18(7):1349-1352
  7. 7. Kohtaro O, Makoto T. Platinum and palladium complexes with metal–silicon bonds. New bonding, structures, and chemical properties. Bulletin of the Chemical Society of Japan. 2005;78(11):1887-1898
  8. 8. Cuevas-Chávez CA, Zamora-Moreno J, Muñoz-Hernández MA, Bijani C, Sabo-Etienne S, Montiel-Palma V. Stabilization of trans disilyl coordination at square-planar platinum complexes. Organometallics, ASAP article, 2017. DOI: 10.1021/acs.organomet.7b00566
  9. 9. Mitton SJ, McDonald R, Turculet L. Synthesis and characterization of neutral and cationic platinum(II) complexes featuring pincer-like Bis(phosphino)silyl ligands: Si−H and Si−Cl bond activation chemistry. Organometallics. 2009;28(17):5122-5136
  10. 10. Shimada S, Tanaka M. Group 10 transition-metal complexes with metal–silicon bonds derived from 1,2-disilylbenzenes and bis(2-silylphenyl)silane. Coordination Chemistry Reviews. 2006;250(9):991-1011
  11. 11. Shimada S, Tanaka M, Honda K. Unusual reactivity of 1,2-disilylbenzene toward Pt(0) complexes. Isolation of the first PtIVSi4P2 and dinuclear, mixed-valence PtIIPtIVSi4P4 complexes. Journal of the American Chemical Society. 1995;117(31):8289-8290
  12. 12. Joslin FL, Stobart SR. (Phosphinoalkyl)silanes. 2. Synthesis and spectroscopic properties of poly(phosphinoalkyl)silanes. Inorganic Chemistry. 1993;32:2221-2223
  13. 13. Mitton SJ, McDonald R, Turculet L. Nickel and palladium silyl pincer complexes: Unusual structural rearrangements that involve reversible Si-C(sp3) and Si-C(sp2) bond activation. Angewandte Chemie (International Edition in English). 2009;48(45):8568-8571
  14. 14. Mitton SJ, McDonald R, Turculet L. Facile intramolecular silicon–carbon bond activation at Pt(0) and Pt(II) centers. Polyhedron. 2013;52:750-754
  15. 15. Takaya J, Iwasawa N. Hydrocarboxylation of allenes with CO2 catalized by silyl pincer-type palladium complexes. Journal of the American Chemical Society. 2008;130:15254-15255
  16. 16. Takaya J, Kirai N, Iwasawa N. Efficient synthesis of diborylalkenes from alkenes and diboron by a new PSiP-pincer palladium-catalyzed dehydrogenative borylation. Journal of the American Chemical Society. 2011;133(33):12980-12983
  17. 17. Kirai N, Iguchi S, Ito T, Takaya J, Iwasawa N. PSiP-pincer type palladium-catalyzed dehydrogenative borylation of alkenes and 1,3-dienes. Bulletin of the Chemical Society of Japan. 2013;86(7):784-799
  18. 18. Takaya J, Iwasawa N. Silyl ligand mediated reversible beta-hydrogen elimination and hydrometalation at palladium. Chemistry—A European Journal. 2014;20(37):11812-11819
  19. 19. Korshin EE, Leitus G, Shimon LJW, Konstantinovski L, Milstein D. Silanol-based pincer Pt(II) complexes: Synthesis, structure, and unusual reactivity. Inorganic Chemistry. 2008;47(16):7177-7189
  20. 20. Morgan E, MacLean DF, McDonald R, Turculet L. Rhodium and iridium amido complexes supported by silyl pincer ligation: Ammonia N-H bond activation by a [PSiP]Ir complex. Journal of the American Chemical Society. 2009;131:14234-14236
  21. 21. Takaya J, Iwasawa N. Bis(o-phosphinophenyl)silane as a scaffold for dynamic behavior of H−Si and C−Si bonds with palladium(0). Organometallics. 2009;28(23):6636-6638
  22. 22. Takaya J, Iwasawa N. Reaction of bis(o-phosphinophenyl)silane with M(PPh3)4 (M = Ni, Pd, Pt): Synthesis and structural analysis of η2-(Si–H) metal(0) and pentacoordinate silyl metal(II) hydride complexes of the Ni triad bearing a PSiP-pincer ligand. Dalton Transactions. 2011;40(35):8814-8821
  23. 23. Joost M, Mallet-Ladeira S, Miqueu K, Amgoune A, Bourissou D. σ-SiH complexes of copper: Experimental evidence and computational analysis. Organometallics. 2013; 32(3):898-902
  24. 24. MacInnis MC, McDonald R, Ferguson MJ, Tobisch S, Turculet L. Four-coordinate, 14-electron Ru(II) complexes: Unusual trigonal pyramidal geometry enforced by bis(phosphino) silyl ligation. Journal of the American Chemical Society. 2011;133(34):13622-13633
  25. 25. Li Y-H, Ding X-H, Zhang Y, He W-R, Huang W. Synthesis, characterization, and catalytic behavior of a PSiP pincer-type ruthenium(II) complex. Inorganic Chemistry Communications. 2012;15:194-197
  26. 26. MacInnis MC, MacLean DF, Lundgren RJ, McDonald R, Turculet L. Synthesis and reactivity of platinum group metal complexes featuring the new pincer-like bis(phosphino) silyl ligand [K3-(2-Ph2PC6H4)2SiMe]-([PSiP]): Application in the ruthenium-mediated transfer hydrogenation of ketones. Organometallics. 2007;26:6522-6525
  27. 27. Kirai N, Takaya J, Iwasawa N. Two reversible sigma-bond metathesis pathways for boron-palladium bond formation: Selective synthesis of isomeric five-coordinate borylpalladium complexes. Journal of the American Chemical Society. 2013;135(7):2493-2496
  28. 28. Takaya J, Kirai N, Iwasawa N. Mechanistic studies on the stereoisomerization between two stereoisomeric, isolable five-coordinate borylpalladium(II) complexes bearing a phenylene-bridged PSiP-pincer type ligand. Organometallics. 2014;33(6):1499-1502
  29. 29. Wu S, Li X, Xiong Z, Xu W, Lu Y, Sun H. Synthesis and reactivity of silyl iron, cobalt, and nickel complexes bearing a [PSiP]-pincer ligand via Si–H bond activation. Organometallics. 2013;32(11):3227-3237
  30. 30. MacInnis MC, Ruddy AJ, McDonald R, Ferguson MJ, Turculet L. Synthesis and characterization of five-coordinate, 16-electron RuII complexes supported by tridentate bis(phosphino)silyl ligation. Dalton Transactions. 2016;45(40):15850-15858
  31. 31. Xu S, Li X, Zhang S, Sun H. Synthesis and characterization of stable tripodal silyl iron and nickel complexes. Inorganica Chimica Acta. 2015;430:161-167
  32. 32. Mankad NP, Whited MT, Peters JC. Terminal Fe(I)-N2 and Fe(II)…H-C interactions supported by tris(phosphino)silyl ligands. Angewandte Chemie (International Edition in English). 2007;46(30):5768-7571
  33. 33. Whited MT, Mankad NP, Lee Y, Oblad PF, Peters JC. Dinitrogen complexes supported by tris(phosphino)silyl ligands. Inorganic Chemistry. 2009;48:2507-2517
  34. 34. Takaoka A, Mendiratta A, Peters JC. E−H bond activation reactions (E = H, C, Si, Ge) at ruthenium: Terminal phosphides, silylenes, and germylenes. Organometallics. 2009; 28(13):3744-3753
  35. 35. Fang H, Choe YK, Li Y, Shimada S. Synthesis, structure, and reactivity of hydridoiridium complexes bearing a pincer-type PSiP ligand. Chemistry, an Asian Journal. 2011; 6(9):2512-2521
  36. 36. Ruddy AJ, Mitton SJ, McDonald R, Turculet L. 'Hemilabile' silyl pincer ligation: Platinum group PSiN complexes and triple C-H activation to form a (PSiC)Ru carbene complex. Chemical communications (Cambridge, England). 2012;48(8):1159-1161
  37. 37. Takaya J, Ito S, Nomoto H, Saito N, Kirai N, Iwasawa N. Fluorine-controlled C-H borylation of arenes catalyzed by a PSiN-pincer platinum complex. Chemical Communications. 2015;51(100):17662-17665
  38. 38. Mankad NP, Müller P, Peters JC. Catalytic N-N coupling of aril azides to yield azoarenes via trigonal bipyramid iron-nitrene intermediates. Journal of the American Chemical Society Communications. 2010;132:4083-4085
  39. 39. Lee Y, Mankad NP, Peters JC. Triggering N2 uptake via redox-induced expulsion of coordinated NH3 and N2 silylation at trigonal bipyramidal iron. Nature Chemistry. 2010;2(7):558-565
  40. 40. Lee Y, Peters JC. Silylation of iron-bound carbon monoxide affords a terminal Fe carbyne. Journal of the American Chemical Society. 2011;133(12):4438-4446
  41. 41. Fong H, Peters JC. Hydricity of an Fe-H species and catalytic CO2 hydrogenation. Inorganic Chemistry. 2015;54(11):5124-5135
  42. 42. Rittle J, Peters JC. Proton-coupled reduction of an iron cyanide complex to methane and ammonia. Angewandte Chemie (International Edition in English). 2016;55:1-5
  43. 43. Rittle J, Peters JC. An Fe-N2 complex that generates hydrazine and ammonia via Fe=NNH2: Demonstrating a hybrid distal-to-alternating pathway for N2 reduction. Journal of the American Chemical Society. 2016;138(12):4243-4248
  44. 44. Mankad NP, Müller P, Peters JC. Catalytic N−N coupling of aryl azides to yield azoarenes via trigonal bipyramid iron−nitrene intermediates. Journal of the American Chemical Society. 2010;132(12):4083-4085
  45. 45. Takaoka A, Peters JC. A homologous series of cobalt, rhodium, and iridium metalloradicals. Inorganic Chemistry. 2012;51(1):16-18
  46. 46. Suess DL, Tsay C, Peters JC. Dihydrogen binding to isostructural S = (1/2) and S = 0 cobalt complexes. Journal of the American Chemical Society. 2012;134(34):14158-14164
  47. 47. Tsay C, Peters JC. Thermally stable N2 and H2 adducts of cationic nickel(II). Chemical Science. 2012;3(4):1313-1318
  48. 48. Takaoka A, Gerber LC, Peters JC. Access to well-defined ruthenium(I) and osmium(I) metalloradicals. Angewandte Chemie (International Edition in English). 2010;49(24):4088-4091
  49. 49. Takaoka A, Mankad NP, Peters JC. Dinitrogen complexes of sulfur-ligated iron. Journal of the American Chemical Society. 2011;133(22):8440-8443
  50. 50. Rittle J, McCrory CCL, Peters JC. A 106-fold enhancement in N2-binding affinity of an Fe2(μ-H)2 core upon reduction to a mixed-valence FeIIFeI state. Journal of the American Chemical Society. 2014;136(39):13853-13862
  51. 51. Connor BA, Rittle J, VanderVelde D, Peters JC. A Ni0(η2-(Si–H))(η2-H2) complex that mediates facile H atom exchange between two σ-ligands. Organometallics. 2016;35(5):686-690
  52. 52. Creutz SE, Peters JC. Diiron bridged-thiolate complexes that bind N2 at the Fe(II)Fe(II), Fe(II)Fe(I), and Fe(I)Fe(I) redox states. Journal of the American Chemical Society. 2015; 137(23):7310-7313
  53. 53. Lee Y-J, Lee J-D, Kim S-J, Keum S, Ko J, Suh I-H, et al. Synthesis, structure, and DFT calculation of (phosphino-o-carboranyl)silyl group 10 metal complexes: Formation of stable trans-Bis(P,Si-chelate)metal complexes. Organometallics. 2004;23(2):203-214
  54. 54. Creutz SE, Peters JC. Exploring secondary-sphere interactions in Fe-N x H y complexes relevant to N2 fixation. Chemical Science. 2017;8(3):2321-2328
  55. 55. Brost RD, Bruce GC, Joslin FL, Stobart SR. Phosphinoalkylsilyl complexes. 12. Stereochemistry of the tridentate Bis(diphenylphosphinopropyl)silyl (biPSi) framework: Complexation that introduces “face discrimination” at Coordinatively unsaturated metal centers. X-ray crystal and molecular structures of Pt[SiMe(CH2CH2CH2PPh2)2]Cl, IrH[SiMe(CH2CH2CH2PPh2)2]Cl, and RuH[SiMe(CH2CH2CH2PPh2)2](CO)2. Organometallics. 1997;16(26):5669-5680
  56. 56. Sola E, Garcia-Camprubi A, Andrés JL, Martin M, Plou P. Iridium compounds with κ-P,P,Si (biPSi) pincer ligands: Favoring reactive structures in unsaturated complexes. Journal of the American Chemical Society. 2010;132:1911-1921
  57. 57. Garcia-Camprubi A, Martin M, Sola E. Addition of water across Si-Ir bonds in iridium complexes with kappa-P,P,Si (biPSi) pincer ligands. Inorganic Chemistry. 2010; 49(22):10649-10657
  58. 58. Montiel-Palma V, Muñoz-Hernandez MA, Ayed T, Barthelat JC, Grellier M, Vendier L, Sabo-Etienne S. Agostic Si-H bond coordination assists C-H bond activation at ruthenium in bis(phosphinobenzylsilane) complexes. Chemical Communications. 2007;0(38):3963-3965
  59. 59. Montiel-Palma V, Muñoz-Hernandez MA, Cuevas-Chavez CA, Vendier L, Grellier M, Sabo-Etienne S. Phosphinodi(benzylsilane) PhP{(o-C6H4CH2)SiMe2H}2: A versatile“PSi2Hx” pincer-type ligand at ruthenium. Inorganic Chemistry. 2013;52(17):9798-9806
  60. 60. Corona-Gonzalez MV, Zamora-Moreno J, Cuevas-Chavez CA, Rufino-Felipe E, Mothes- Martin E, Coppel Y, Muñoz-Hernandez MA, Vendier L, Flores-Alamo M, Grellier M, Sabo-Etienne S, Montiel-Palma V. A family of rhodium and iridium complexes with semirigid benzylsilyl phosphines: From bidentate to tetradentate coordination modes. Dalton Transactions. 2017;46(27):8827-8838
  61. 61. Shimizu H, Nagasaki I, Matsumura K, Sayo N, Saito T. Developments in asymmetric hydrogenation from an industrial perspective. Accounts of Chemical Research. 2007; 40(12):1385-1393
  62. 62. Colacot T. Nobel Prize in Chemistry. Timely recognition for Rh, Ru and Os-catalysed chiral reactions. Platinum Metals Review. 2001, 2002;46(2):82-83
  63. 63. Ikariya T, Murata K, Noyori R. Bifunctional transition metal-based molecular catalysts for asymmetric syntheses. Organic & Biomolecular Chemistry. 2006;4(3):393-406
  64. 64. Tamao K, Sumitani K, Kumada M. Selective carbon-carbon bond formation by crosscoupling of Grignard reagents with organic halides. Catalysis by nickel-phosphine complexes. Journal of the American Chemical Society. 1972;94(12):4374-4376
  65. 65. Hua X, Masson-Makdissi J, Sullivan RJ, Newman SG. Inherent vs apparent chemoselectivity in the Kumada–Corriu cross-coupling reaction. Organic Letters. 2016; 18(20):5312-5315
  66. 66. Yoshikai N, Matsuda H, Nakamura E. Hydroxyphosphine ligand for nickel-catalyzed cross-coupling through nickel/magnesium bimetallic cooperation. Journal of the American Chemical Society. 2009;131(27):9590-9599
  67. 67. Liu N, Wang Z-X. Kumada coupling of aryl, heteroaryl, and vinyl chlorides catalyzed by amido pincer nickel complexes. The Journal of Organic Chemistry. 2011; 76(24):10031-10038
  68. 68. Xi Z, Liu B, Chen W. Room-temperature Kumada cross-coupling of unactivated aryl chlorides catalyzed by N-heterocylic carbene-based nickel(II) complexes. The Journal of Organic Chemistry. 2008;73(10):3954-3957
  69. 69. Iglesias MJ, Prieto A, Nicasio MC. Kumada–Tamao–Corriu coupling of heteroaromatic chlorides and aryl ethers catalyzed by (IPr)Ni(allyl)Cl. Organic Letters. 2012; 14(17):4318-4321
  70. 70. Xiong Z, Li X, Zhang S, Shi Y, Sun H. Synthesis and reactivity of N-heterocyclic PSiP pincer iron and cobalt complexes and catalytic application of cobalt hydride in Kumada coupling reactions. Organometallics. 2016;35(3):357-363

Written By

Julio Zamora-Moreno and Virginia Montiel-Palma

Submitted: 25 September 2017 Reviewed: 04 January 2018 Published: 23 May 2018