首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The reactions of bis(borohydride) complexes [(RN?)Mo(BH4)2(PMe3)2] ( 4 : R=2,6‐Me2C6H3; 5 : R=2,6‐iPr2C6H3) with hydrosilanes afford new silyl hydride derivatives [(RN?)Mo(H)(SiR′3)(PMe3)3] ( 3 : R=Ar, R′3=H2Ph; 8 : R=Ar′, R′3=H2Ph; 9 : R=Ar, R′3=(OEt)3; 10 : R=Ar, R′3=HMePh). These compounds can also be conveniently prepared by reacting [(RN?)Mo(H)(Cl)(PMe3)3] with one equivalent of LiBH4 in the presence of a silane. Complex 3 undergoes intramolecular and intermolecular phosphine exchange, as well as exchange between the silyl ligand and the free silane. Kinetic and DFT studies show that the intermolecular phosphine exchange occurs through the predissociation of a PMe3 group, which, surprisingly, is facilitated by the silane. The intramolecular exchange proceeds through a new non‐Bailar‐twist pathway. The silyl/silane exchange proceeds through an unusual MoVI intermediate, [(ArN?)Mo(H)2(SiH2Ph)2(PMe3)2] ( 19 ). Complex 3 was found to be the catalyst of a variety of hydrosilylation reactions of carbonyl compounds (aldehydes and ketones) and nitriles, as well as of silane alcoholysis. Stoichiometric mechanistic studies of the hydrosilylation of acetone, supported by DFT calculations, suggest the operation of an unexpected mechanism, in that the silyl ligand of compound 3 plays an unusual role as a spectator ligand. The addition of acetone to compound 3 leads to the formation of [trans‐(ArN)Mo(OiPr)(SiH2Ph)(PMe3)2] ( 18 ). This latter species does not undergo the elimination of a Si? O group (which corresponds to the conventional Ojima′s mechanism of hydrosilylation). Rather, complex 18 undergoes unusual reversible β‐CH activation of the isopropoxy ligand. In the hydrosilylation of benzaldehyde, the reaction proceeds through the formation of a new intermediate bis(benzaldehyde) adduct, [(ArN?)Mo(η2‐PhC(O)H)2(PMe3)], which reacts further with hydrosilane through a η1‐silane complex, as studied by DFT calculations.  相似文献   

2.
A 1:1:1 reaction between complex (Tp)(ArN═)Mo(H)(PMe(3)) (3), silane PhSiD(3), and carbonyl substrate established that hydrosilylation catalyzed by 3 is not accompanied by deuterium incorporation into the hydride position of the catalyst, thus ruling out the conventional hydride mechanism based on carbonyl insertion into the M-H bond. An analogous result was observed for the catalysis by (O═)(PhMe(2)SiO)Re(PPh(3))(2)(I)(H) and (Ph(3)PCuH)(6).  相似文献   

3.
The compounds [MoCl(NAr)2R] (R=CH2CMe2Ph (1) or CH2CMe3(2); Ar=2,6-Pri2C6H3) have been prepared from [MoCl2(NAr)2(dme)] (dme=1,2-dimethoxyethane) and one equivalent of the respective Grignard reagent RMgCl in diethyl ether. Similarly, the mixed-imido complex [MoCl2(NAr)(NBut)(dme)] affords [MoCl(NAr)(NBut)(CH2CMe2Ph)] (3). Chloride substitution reactions of 1 with the appropriate lithium reagents afford the compounds [MoCp(NAr)2(CH2CMe2Ph)] (4) (Cp=cyclopentadienyl), [MoInd(NAr)2(CH2CMe2Ph)] (5) (Ind=Indenyl), [Mo(OBut)(NAr)2(CH2CMe 2Ph)] (6), [MoMe(NAr)2(CH2CMe2Ph)] (7), [MoMe(PMe3)(NAr)2(CH2CMe 2Ph)] (8) (formed in the presence of PMe3) and [Mo(NHAr)(NAr)2(CH2CMe2P h)](9). In the latter case, a by-product {[Mo(NAr)2(CH2CMe2Ph) ]2(μ-O)}(10) has also been isolated. The crystal structures of 1, 4, 5 and 10 have been determined. All possess distorted tetrahedral metal centres with cis near-linear arylimido ligands; in each case (except 5, for which the evidence is unclear) there are α-agostic interactions present.  相似文献   

4.
Hydrocarbon solutions of Mo2(O—t-Bu)6 and PF3 (2 equiv) yield Mo4F4(O—t-Bu)8, I, and PF2(O—t-Bu). Compound I contains a bisphenoid of molybdenum atoms with two short MoMo distances, 2.26 Å, and four long MoMo distances, 3.75 Å, corresponding to localized MoMo triple bonding and non-bonding distances, respectively. The tetranuclear compound may be viewed as a dimer, [Mo22-F)2(O-t-Bu)4]2, and addition of PMe3 to hydrocarbon solutions of I yields Mo2F2(O—t-Bu)4(PMe3)2, II, which contains an unbridged MoMo triple bond of distance 2.27 Å. Each molybdenum atom is coordinated to two oxygen atoms, one fluorine atom and the phosphorus atom of the PMe3 ligand in a roughly square planar manner. The overall central Mo2O4F2P2 skeleton has C2 symmetry and NMR studies (1H, 19F and 31P) are consistent with the maintenance of this type of structure in solution. Infrared and electronic absorption spectral data are reported. These are the first compounds containing fluorine ligands attached to the (MoMo)6+ unit.  相似文献   

5.
3,3′,5,5′-Tetra-tert-butyl-2′-sulfanyl[1,1′-biphenyl]-2-ol (H2[tBu4OS]) was prepared in 24 % yield overall from the analogous biphenol using standard techniques. Addition of H2[tBu4OS] to Mo(NAr)(CHCMe2Ph)(2,5-dimethylpyrrolide)2 led to formation of Mo(NAr)(CHCMe2Ph)[tBu4OS], which was trapped with PMe3 to give Mo(NAr)(CHCMe2Ph)[tBu4OS](PMe3) ( 1 (PMe3)). An X-ray crystallographic study of 1 (PMe3) revealed that two structurally distinct square pyramidal molecules are present in which the alkylidene ligand occupies the apical position in each. Both 1 (PMe3)A and 1 (PMe3)B are disordered. Mo(NAd)(CHCMe2Ph)(tBu4OS)(PMe3) ( 2 (PMe3); Ad=1-adamantyl) and W(NAr)(CHCMe2Ph)(tBu4OS)(PMe3) ( 3 (PMe3)) were prepared using analogous approaches. 1 (PMe3) reacts with ethylene (1 atm) in benzene within 45 minutes to give an ethylene complex Mo(NAr)(tBu4OS)(C2H4) ( 4 ) that is isolable and relatively stable toward loss of ethylene below 60 °C. An X-ray study shows that the bond distances and angles for the ethylene ligand in 4 are like those found for bisalkoxide ethylene complexes of the same general type. Complex 1 (PMe3) in the presence of one equivalent of B(C6F5)3 catalyzes the homocoupling of 1-decene, allyltrimethylsilane, and allylboronic acid pinacol ester at ambient temperature. 1 (PMe3), 2 (PMe3), and 3 (PMe3) all catalyze the ROMP of rac-endo,exo-5,6-dicarbomethoxynorbornene (rac-DCMNBE) in the presence of B(C6F5)3, but the polyDCMNBE that is formed has a random structure.  相似文献   

6.
The feasibility of oxidative addition of the P−H bond of PHPh2 to a series of rhodium complexes to give mononuclear hydrido-phosphanido complexes has been analyzed. Three main scenarios have been found depending on the nature of the L ligand added to [Rh(Tp)(C2H4)(PHPh2)] (Tp= hydridotris(pyrazolyl)borate): i) clean and quantitative reactions to terminal hydrido-phosphanido complexes [RhTp(H)(PPh2)(L)] (L=PMe3, PMe2Ph and PHPh2), ii) equilibria between RhI and RhIII species: [RhTp(H)(PPh2)(L)]⇄[RhTp(PHPh2)(L)] (L=PMePh2, PPh3) and iii) a simple ethylene replacement to give the rhodium(I) complexes [Rh(κ2-Tp)(L)(PHPh2)] (L=NHCs-type ligands). The position of the P−H oxidative addition–reductive elimination equilibrium is mainly determined by sterics influencing the entropy contribution of the reaction. When ethylene was used as a ligand, the unique rhodaphosphacyclobutane complex [Rh(Tp)(η1-Et)(κC,P-CH2CH2PPh2)] was obtained. DFT calculations revealed that the reaction proceeds through the rate limiting oxidative addition of the P−H bond, followed by a low-barrier sequence of reaction steps involving ethylene insertion into the Rh−H and Rh−P bonds. In addition, oxidative addition of the P−H bond in OPHPh2 to [Rh(Tp)(C2H4)(PHPh2)] gave the related hydride complex [RhTp(H)(PHPh2)(POPh2)], but ethyl complexes resulted from hydride insertion into the Rh−ethylene bond in the reaction with [Rh(Tp)(C2H4)2].  相似文献   

7.
Photoreaction of diaminosubstituted-phosphiteborane, BH3P(NMeCH2)2(OMe) with a methyl molybdenum complex, (η5-C5R5)Mo(CO)3Me (R5 = Me5, Me4H, H5) yielded a phosphiteboryl molybdenum complex, (η5-C5R5)Mo(CO)3BH2{P(NMeCH2)2(OMe)} (R5 = Me5: 2, Me4H: 3, H5: 4). In the reaction of 2 with MeI, the Mo–B bond was activated to give (η5-C5Me5)Mo(CO)3Me, in the reaction with PMe3, the B–P bond was activated to give (η5-C5Me5)Mo(CO)3(BH2PMe3). Complex 2 in solution was gradually converted into (η5-C5Me5)MoH(CO)2{P(NMeCH2)2(OMe)} (8) via the B–H bond activation of 2. Structures of 2, 3, and 8 were determined by single crystal X-ray diffraction studies.  相似文献   

8.
The heteroelement-containing alkylidene imide complexes with molybdenum and tungsten Et3SiCH=Mo(NAr)(OR)2 (I), Et3 ECH=W(NAr)(OR)2 (E = Si (II), Ge (III); Ar = 2,6-i-Pr2C6H3; R=CMe2 CF3) and π-complex (RO)2(ArN)Mo(CH2=CH-GeEt3) (IV) were synthesized by the reaction of Alkyl-CH=M(NAr) (OR)2 (M=Mo, W; Alkyl = t-Bu, PhMe2C) with organosilicon and organogermanium vinyl reagents Et3ECH=CH2 (E = Si, Ge). The structure of compounds I–III was determined by X-ray diffraction (XRD). The complexes I–IV are active initiators of metathesis polymerization of cycloolefins.  相似文献   

9.
The cyclopentadienyl molybdenum hydride compounds, CpRMo(PMe3)3–x(CO)xH (CpR = Cp, Cp*; x = 0, 1, 2 or 3), are catalysts for the dehydrogenation of formic acid, with the most active catalysts having the composition CpRMo(PMe3)2(CO)H. The mechanism of the catalytic cycle is proposed to involve (i) protonation of the molybdenum hydride complex, (ii) elimination of H2 and coordination of formate, and (iii) decarboxylation of the formate ligand to regenerate the hydride species. NMR spectroscopy indicates that the nature of the resting state depends on the composition of the catalyst. For example, (i) the resting states for the CpMo(CO)3H and CpMo(PMe3)(CO)2H systems are the hydride complexes themselves, (ii) the resting state for the CpMo(PMe3)3H system is the protonated species [CpMo(PMe3)3H2]+, and (iii) the resting state for the CpMo(PMe3)2(CO)H system is the formate complex, CpMo(PMe3)2(CO)(κ1-O2CH), in the presence of a high concentration of formic acid, but CpMo(PMe3)2(CO)H when the concentration of acid is low. While CO2 and H2 are the principal products of the catalytic reaction induced by CpRMo(PMe3)3–x(CO)xH, methanol and methyl formate are also observed. The generation of methanol is a consequence of disproportionation of formic acid, while methyl formate is a product of subsequent esterification. The disproportionation of formic acid is a manifestation of a transfer hydrogenation reaction, which may also be applied to the reduction of aldehydes and ketones. Thus, CpMo(CO)3H also catalyzes the reduction of a variety of ketones and aldehydes to alcohols by formic acid, via a mechanism that involves ionic hydrogenation.  相似文献   

10.
Special reaction conditions enable dicarbonyl(η5-cyclopentadienyl)tolylcarbyne complexes of molybdenum and tungsten to react with trimethylphosphine by the replacement of one carbonyl ligand to give Cp(CO)(PMe3WC-Tol (M = Mo, W). The preparation and spectroscopic data of these complexes as well as structural parameters of the tungsten complex are reported.  相似文献   

11.
The platinum trimethyl complex of a benzoic acid-functionalized hydrido-tris(pyrazolyl)borate (Tp) ligand [p-(HO2C)C6H4Tp]PtMe34 has been synthesized from the corresponding p-bromo complex [p-BrC6H4Tp]PtMe33. Compound 4 may be readily coupled to biomolecules such as amino acids as exemplified by coupling to l-phenylalanine-tert-butylester to provide [p-(tBuO-Phe-CO)C6H4Tp]PtMe35. Compound 5 has been structurally characterized in the solid state by X-ray diffraction. It constitutes the first example of a tris(pyrazolyl)borate bioconjugate.  相似文献   

12.
Green and blue isomers of the oxo derivative MoOCl2(PMe3)3 have been obtained by an oxygen-atom abstraction reaction between MoCl4(thf)2 and equimolar amounts of water in the presence of PMe3. Methatesis with KX(X = NCO, NCS) yields MoOX2(PMe3)3 and with NaS2CNEt2, MoO(S2CNEt2)2(PMe3). The latter complex readily loses PMe3 to give MoO(S2CNEt2)2 from which it can be prepared by addition of the phosphine ligand.Reaction of the blue purple complex, MoCl3(thf)3 (I), with excess PMe3 gives mer-MoCl3(PMe3)3 which loses PMe3 on heating in toluene to afford [MoCl3(PMe3)2]2. Reduction of (I) with phosphines and zinc in tetrahydrofuran gives the dinuclear molybdenum(II) halide complexes Mo2Cl4L4 (L = PMe3, PEt3, PhMe2Ph, PEt2Ph; L2 = dppm), while Zn-acetic acid reduction yields Mo2(CO2Me)4. Interaction of the chlorocarbonyl species MoCl2(CO)2(PMe3)3 with Tl(acac) affords Mo(acac)Cl(CO)(PMe3)3 which has an unusually low CO stretching frequency for a terminal carbonyl group (1755 cm?1).  相似文献   

13.
Reaction of [W(PMe2Ph)3H6] with pentaborane(9) gives nido-2-[W(PMe2Ph)3H2B4H8] (1) as well as nido-2-[W(PMe2Ph)3HB5H10] (2). The crystal structure of (2) has been determined. Compound (2) has a novel metallaborane structure containing an edge-bridging {BH3} group between the tungsten atom and one of the basal boron atoms in a “nido-WB4” pyramid. Reaction of [W(PMe3)42-CH2PMe2)H] with pentaborane(9) gives nido-2-[W(PMe3)3H2B4H8] (3) whilst reaction of [Mo(L)4H4] with pentaborane(9) gives nido-2-[Mo(L)3H2B4H8] [L = PMe3 (4), PMe2Ph (5)]. Treatment of [Mo(PMe3)4H4] with excess BH3 · thf gives the known borohydride [Mo(PMe3)4H(η2-BH4)].  相似文献   

14.
A very effective solid support for the removal of molybdenum-based metathesis catalysts can be prepared by placing a salicylimine on a polystyrene support. The resin is produced by treating Merrifield’s resin with H2NBun, 4-chloromethylsalicylaldehyde, and p-toluidine in succession. The final resin had an identifiable CN stretch in the IR and a resonance assigned to the HCN hydrogen was found by MAS 1H NMR of the swelled resin. Solutions of Mo[C(H)CMe2Ph](NAr)(OButF6)2 (MoF6) and Mo[C(H)CMe2Ph](NAr)(OAd)2 (MoAd), where Ar = 2,6-diisopropylphenyl and Ad = 1-adamantyl, were treated with the scavenger, which reduced the remaining molybdenum concentration as examined by ICP-MS to 30-50 ppb. Catalyst was also scavenged from ring-closing metathesis of diethyl diallylmalonate by MoF6; the decrease in molybdenum concentration on addition of scavenger followed first order kinetics with initial and final concentrations of 54 000 and 15 ppb, respectively. We also prepared a model system where a soluble salicylimine (H-DIB) was reacted with Mo[C(H)CMe2Ph](NAr)(OAd)2 to produce Mo[C(H)CMe2Ph](NAr)(DIB)2, which was structurally characterized.  相似文献   

15.
The reaction of the dimeric bis(germylene) [Ge{3,5‐(CF3)2pz}2]2 ( 2 ) with protic molybdenum hydride [Mo(H)Cp(CO)3] yielded two different products. In diethyl ether the divalent germylene readily inserts into the Mo–H σ‐bond and the product of the oxidative addition, [Ge(H){Mo}(pz)2] ( 4 ) (with pz = 3,5‐disubstituted pyrazole, 3,5‐(CF3)2pz; {Mo} = [MoCp(CO)3]), was isolated featuring a germanium(IV) hydride moiety. In toluene an interesting “cascade” reaction takes place furnishing a bis‐metal substituted digermane [{Mo}(H)(pz)Ge–Ge(pz)2{Mo}]. Although the detailed mechanism of the reaction remains the subject of speculation it seems likely that a germylene, [GeII(pz){Mo}], inserts into the germanium(IV) hydrogen bond of [Ge(H){Mo}(pz)2] under formation of a germanium‐germanium bond, which is a rare reaction behaviour.  相似文献   

16.
The keto-functionalised N-pyrrolyl phosphine ligand PPh2NC4H3{C(O)CH3-2} L1 reacts with [MoCl(CO)35-C5R5)] (R=H, Me) to give [MoCl(CO)2(L11P)(η5-C5R5)] (R=H 1a; Me 1b). The phosphine ligands PPh2CH2C(O)Ph (L2) and PPh2CH2C(O)NPh2 (L3) react with [MoCl(CO)35-C5R5)] in an analogous manner to give the compounds [MoCl(CO)2(L-κ1P)(η5-C5R5)] (L=L2, R=H 2a, Me 2b; L=L3, R=H 3a, Me 3b). Compounds 13 react with AgBF4 to give [Mo(CO)2(L-κ2P,O)(η5-C5R5)]BF4 (L=L1, R=H 4a, Me 4b; L=L2, R=H 5a, Me 5b; L=L3, R=H 6a, Me 6b) following displacement of chloride. The X-ray crystal structure of 4a revealed a lengthening of both Mo–P and CO bonds on co-ordination of the keto group. The lability of the co-ordinated keto or amido group has been assessed by addition of a range of phosphines to compounds 46. Compound 4a reacts with PMe3, PMe2Ph and PMePh2 to give [Mo(CO)2(L11P)(L)(η5-C5H5)]BF4 (L=PMe3 7a; PMe2Ph 7b; PMePh2 7c) but does not react with PPh3, 5a reacts with PMe2Ph, PMePh2 and PPh3 to give [Mo(CO)2(L21P)(L)(η5-C5H5)]BF4 (L=PMe2Ph 8b; PMePh2 8c; PPh3 8d), and 6a reacts with PMe3, PMe2Ph, PMePh2 and PPh3 to give [Mo(CO)2(L31P)(L)(η5-C5H5)]BF4 (L=PMe3 10a; PMe2Ph 10b; PMePh2 10c; PPh3 10d). No reaction was observed for the pentamethylcyclopentadienyl compounds 4b6b with PMe3, PMe2Ph, PMePh2 or PPh3. These results are consistent with the displacement of the co-ordinated oxygen atom being influenced by the steric properties of the P,O-ligand, with PPh3 displacing the keto group from L2 but not from the bulkier L1. In the reaction of [Mo(CO)2(L22P,O)(η5-C5H5)]BF4 (5a) with PMe3 the phosphine does not displace the keto group, instead it acts as a base, with the only observed molybdenum-containing product being the enolate compound [Mo(CO)2{PPh2CHC(O)Ph-κ2P,O}(η5-C5H5)] 9. Compound 9 can also be formed from the reaction of 2a with BuLi or NEt3, and a single crystal X-ray analysis has confirmed the enolate structure.  相似文献   

17.
The treatment of the hexacarbonylmetal compounds M(CO)6 (M = Cr. Mo, W) with two equivalents Me3PCH2 yields the phosphonium acylmetalphosphorus ylides Me4P[(CO)5MC(O)CHPMe3] 1a–1c. Their reaction with Me3SiOSO2CF3 leads via O-silylation to formation of the neutral “siloxy(ylidecarbene) complexes” (CO)5MC(OSiMe3)CHPMe32a–2c, which are protonated by HX (X = Cl, CF3SO3) to give the thermolabile carbene complexes [(CO)5MC(OSiMe3)H2CPMe3]X, 3a, 3b. 1H, 13C NMR and IR data suggest, that delocalization of the ylidic charge to the carbene carbon generates a metal-coordinated vinyl group in the case of 2a–2c. In addition this fact is proved by the X-ray analysis of 2c, for which a C(ylide)C(carbene) bond distance of 133 pm is found. 2a–2c are obtained as pure E-isomers but can be converted to the Z-isomers 2a′–2c′ upon photolysis.  相似文献   

18.
Relativistic density functional calculations have been carried out for the group VI transition metal carbonyls M(CO)5L (M=Cr, Mo, W; L=OH2, NH3, PH3, PMe3, N2, CO, OC (isocarbonyl), CS, CH2, CF2, CCl2, NO+). The optimized molecular structures and M(SINGLE BOND)L bond dissociation energies, as well as the metal–carbonyl bond energy of the trans CO group, have been calculated. Besides the marked dependence of the trans M(SINGLE BOND)CO bond length on the type of ligand L, such an effect on the that bond energy is also observed. For the chromium compounds, the trans Cr(SINGLE BOND)CO bond length varies from 184 to 199 pm and its bond energy from 242 to 150 kJ/mol. For the molybdenum compounds, the range is 197 to 216 pm and 253 to 128 kJ/mol and, for tungsten, 198 to 214 pm and 293 to 159 kJ/mol. The observed trends can be explained with the π acceptor strength of the L ligand. © 1997 John Wiley & Sons, Inc. J Comput Chem 18 : 1985–1992, 1997  相似文献   

19.
Described herein is the development of the B(C6F5)3‐catalyzed hydrosilylation of α,β‐unsaturated esters and amides to afford synthetically valuable α‐silyl carbonyl products. The α‐silylation occurs chemoselectively, thus leaving the labile carbonyl groups intact. The reaction features a broad scope of both acyclic and cyclic substrates, and the synthetic utility of the obtained α‐silyl carbonyl products is also demonstrated. Mechanistic studies revealed two operative steps: fast 1,4‐hydrosilylation of conjugated carbonyls and then slow silyl group migration of a silyl ether intermediate.  相似文献   

20.
Treatment of LiC(SiMe2H)3]·2THF (1) with alkeny1chlorosilanes produced sterically hindered alkenylsilanes (410) of structure H2C=CH---(CH2)nSiRR′C(SiMe2H)3 (R=Me; R′=Me or Cl; n=0, 1, or 4). The Peterson reaction of 1 with carbonyl compounds gave sterically hindered olefins R(R′)C=C(SiMe2H)2. Pt or Rh catalyzed intramolecular hydrosilylation of H2C=CHSiMe2C(SiMe2H)3 (4) occurred to produce a new 1,3-disilacyclobutane derivative 15. Intermolecular hydrosilylation was favored for 5, 8, and 10, producing oligomeric products.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号