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 873 毫秒
1.
Self‐immobilized nickel and iron diimine catalysts bearing one or two allyl groups of [ArN?C]2(C10H6)NiBr2 [Ar = 4‐allyl‐2,6‐(i‐Pr)2C6H2] ( 1 ), [ArN?C(Me)][Ar′N? C(Me)]C5H3NFeCl2 [Ar = Ar′ = 4‐allyl‐2,6‐(i‐Pr)2C6H3, Ar = 2,6‐(i‐Pr)2C6H3, and Ar′ = 4‐allyl‐2,6‐(i‐Pr)2C6H3] were synthesized and characterized. All three catalysts were investigated for olefin polymerization. As a result, these catalysts not only showed high activities as the catalyst free from the allyl group, such as [ArN?C]2C10H6NiBr2 (Ar = 2,6‐(i‐Pr)2C6H2)], but also greatly improved the morphology of polymer particles to afford micron‐granula polyolefin. The self‐immobilization of catalysts, the formation mechanism of microspherical polymer, and the influence on the size of the particles are discussed. The molecular structure of self‐immobilized nickel catalyst 1 was also characterized by crystallographic analysis. © 2004 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 42: 1018–1024, 2004  相似文献   

2.
Norbornene polymerizations were carried out using nickel(II) bromide complexes CH{C(R)NAr}2NiBr ( 1 , R = CH3, Ar = 2, 6 ? iPr2C6H3; 2 , R = CH3, Ar = 2, 6‐Me2C6H3; 3 , R = CF3, Ar = 2, 6 ? iPr2C6H3; 4 , R = CF3, Ar = 2, 6‐Me2C6H3) in the presence of methylaluminoxane. Compound 3 is the most active norbornene polymerization catalyst of all the nickel complexes tested. The activity of theses catalysts increases with increases in steric bulk of the substituents on the aryl rings. The electronic nature of the ligand backbone also affects the activity. The resulting polynorbornenes are vinyl type by IR and NMR analyses. Copyright © 2007 John Wiley & Sons, Ltd.  相似文献   

3.
Polymerization of styrene using β‐diketiminate nickel (II) bromide complexes CH{C(R)NAr}2NiBr (R = CH3, Ar = 2,6‐iPr2C6H3, 1 ; R = CH3, Ar = 2,6‐Me2C6H3, 2 ; R = CF3, Ar = 2,6‐iPr2C6H3, 3 ; R = CF3, Ar = 2,6‐Me2C6H3, 4 ) in the presence of methylaluminoxane was studied. Compound 3 is the most active styrene polymerization catalyst of all the nickel complexes tested. The activity of these catalysts increases with increases in steric bulk of the substituents on the aryl rings. The electronic nature of the ligand backbone also affects the activity. Weight‐average molecular weight of the prepared polystyrene ranges from 21 000 to 72 000, with polydispersity indexes of 1.95–2.78. The microstructure of the obtained products is atactic polystyrenes from NMR analyses. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

4.
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   

5.
A series of salicylaldimine‐based neutral Ni(II) complexes (3a–j) [ArN = CH(C6H4O)]Ni(PPh3)Ph [3a, Ar = C6H5; 3b, Ar = C6H4F(o); 3c, Ar = C6H4F(m); 3d, Ar = C6H4F(p); 3e, Ar = C6H3F2(2,4); 3f, Ar = C6H3F2(2,5); 3g, Ar = C6H3F2(2,6); 3h, Ar = C6H3F2(3,5); 3i, Ar = C6H2F3(3,4,5); 3j, Ar = C6F5] have been synthesized in good yield, and the structures of complexes 3a and 3i have been confirmed by X‐ray crystallographic analysis. Using modified methylaluminoxane as a cocatalyst, these neutral Ni(II) complexes exhibited high catalytic activities for the vinylic polymerization of norbornene. It was observed that the strong electron‐withdrawing effect of the fluorinated salicylaldiminato ligand was able to significantly increase the catalyst activity for vinylic polymerization of norbornenes. In addition, catalyst activity, polymer yield and polymer molecular weight can also be controlled over a wide range by the variation of reaction parameters such as Al:Ni ratio, norbornene:catalyst ratio, monomer concentration, polymerization temperature and time. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

6.
The first α‐diimine nickel(I) complex having a chloro bridge is reported. The centrosymmetric dinuclear structure of {[ArN?C(Me)C(Me)?NAr]NiCl}2[Ar?2,6?C6H3(i‐Pr)2] features two chelating α‐diimine ligands and two bridged chlorine atoms, so that a distorted tetrahedral N2Cl2 coordination geometry for nickel results. Copyright © 2004 John Wiley & Sons, Ltd.  相似文献   

7.
Three isothiocyanate complexes of nickel(II) containing diimine [ArN?C(Me)? C(Me)?NAr]Ni‐ (NCS)2 (1), iminophosphine [Ph2PC6H4CH?NAr]Ni(NCS)2 (2), or diphosphine (dppe)Ni(NCS)2 (3), [Ar = 2, 6‐iPr‐C6H3; dppe = 1, 2‐bis(diphenylphosphine)ethane] were synthesized and examined for ethylene polymerization activated by methylaluminoxane (MAO). Their behavior was compared with those of the corresponding halide analogues [ArN?C(Me)? C(Me)?NAr]NiBr2 (4), [Ph2PC6H4CH?NAr]NiBr2 (5), and (dppe)NiCl2 (6). The diimines showed the highest polymerization activity. Replacement of the halide for the NCS pseudo halide affected the activity and decreased the molecular weight of the polymer formed. The highest molecular weights were obtained with the diimine complexes. Highly branched polyethylenes were obtained with the bulkier complexes 1 and 4. Replacement of the halide for NCS in the diimine complexes also caused an increase in the branching content, whereas the opposite occurs for the iminophosphine complexes. The different activities and behavior of the catalyst systems with halide versus NCS in the polymerization of ethylene and the characteristics of the final products suggest a modification in the active species caused by the non‐chelating ligand. Polymer molecular weight and branching content is dependent on the MAO/Ni molar ratio and on the working temperature. Copyright © 2004 John Wiley & Sons, Ltd.  相似文献   

8.
The reaction of [(ArN)2MoCl2] · DME (Ar = 2,6‐i‐Pr2C6H3) ( 1 ) with lithium amidinates or guanidinates resulted in molybdenum(VI) complexes [(ArN)2MoCl{N(R1)C(R2)N(R1)}] (R1 = Cy (cyclohexyl), R2 = Me ( 2 ); R1 = Cy, R2 = N(i‐Pr)2 ( 3 ); R1 = Cy, R2 = N(SiMe3)2 ( 4 ); R1 = SiMe3, R2 = C6H5 ( 5 )) with five coordinated molybdenum atoms. Methylation of these compounds was exemplified by the reactions of 2 and 3 with MeLi affording the corresponding methylates [(ArN)2MoMe{N(R1)C(R2)N(R1)}] (R1 = Cy, R2 = Me ( 6 ); R1 = Cy, R2 = N(i‐Pr)2 ( 7 )). The analogous reaction of 1 with bulky [N(SiMe3)C(C6H5)C(SiMe3)2]Li · THF did not give the corresponding metathesis product, but a Schiff base adduct [(ArN)2MoCl2] · [NH=C(C6H5)CH(SiMe3)2] ( 8 ) in low yield. The molecular structures of 7 and 8 are established by the X‐ray single crystal structural analysis.  相似文献   

9.
Six new arenetelluronic triorganotin esters, namely (R3Sn)4[ArTe(μ‐O)(OH)O2)]2 (Ar = Ph, R = Me: 1 , R = Ph: 2 ; Ar = 3‐Me‐Ph, R = Me: 3 , R = Ph: 4 , Ar = 3‐Cl‐Ph, R = Me: 5 , R = Ph: 6 ), were prepared by treating arenetelluronic acids with the corresponding R3SnCl (R = Me, Ph) with potassium hydroxide in methanol. All complexes were characterized by elemental analysis, FT‐IR, NMR (1H, 13C, 119Sn) spectroscopy, and X‐ray crystallography. The structural analyses indicate that these complexes are isostructural as Sn4Te2 moiety, in which the Te22‐O)2 units are situated in the center and each Te atom is coordinated with two OSnR3 groups on the side. Complexes 1 , 3 , and 5 show one‐dimensional chain and two‐dimensional network supramolecular structures by intermolecular C H···O or C H···Cl interactions. The antitumor activities of these complexes reveal that most arenetelluronic triorganotin esters have powerful antitumor activities with certain regularity.  相似文献   

10.
A series of ruthenium alkenylacetylide complexes trans-[Ru{C≡CC(=CH2)R}Cl(dppe)2] (R=Ph ( 1 a ), cC4H3S ( 1 b ), 4-MeS-C6H4 ( 1 c ), 3,3-dimethyl-2,3-dihydrobenzo[b]thiophene (DMBT) ( 1 d )) or trans-[Ru{C≡C-cC6H9}Cl(dppe)2] ( 1 e ) were allowed to react with the corresponding propargylic alcohol HC≡CC(Me)R(OH) (R=Ph ( A ), cC4H3S ( B ), 4-MeS-C6H4 ( C ), DMBT ( D ) or HC≡C-cC6H10(OH) ( E ) in the presence of TlBF4 and DBU to presumably give alkenylacetylide/allenylidene intermediates trans-[Ru{C≡CC(=CH2)R}{C=C=C(Me)}(dppe)2]PF6 ([ 2 ]PF6). These complexes were not isolated but deprotonated to give the isolable bis(alkenylacetylide) complexes trans-[Ru{C≡CC(=CH2)R}2(dppe)2] (R=Ph ( 3 a ), cC4H3S ( 3 b ), 4-MeS-C6H4 ( 3 c ), DMBT ( 3 d )) and trans-[Ru{C≡C-cC6H9}2(dppe)2] ( 3 e ). Analogous reactions of trans-[Ru(CH3)2(dmpe)2], featuring the more electron-donating 1,2-bis(dimethylphosphino)ethane (dmpe) ancillary ligands, with the propargylic alcohols A or C and NH4PF6 in methanol allowed isolation of the intermediate mixed alkenylacetylide/allenylidene complexes trans-[Ru{C≡CC(=CH2)R}{C=C=C(Me)}(dmpe)2]PF6 (R=Ph ([ 4 a ]PF6), 4-MeS-C6H4 ([ 4 c ]PF6). Deprotonation of [ 4 a ]PF6 or [ 4 c ]PF6 gave the symmetric bis(alkenylacetylide) complexes trans-[Ru{C≡CC(=CH2)R}2(dmpe)2] (R=Ph ( 5 a ), 4-MeS-C6H4 ( 5 c )), the first of their kind containing the dmpe ancillary ligand sphere. Attempts to isolate bis(allenylidene) complexes [Ru{C=C=C(Me)R}2(PP)2]2+ (PP=dppe, dmpe) from treatment of the bis(alkenylacetylide) species 3 or 5 with HBF4 ⋅ Et2O were ultimately unsuccessful.  相似文献   

11.
A series of NCO/NCS pincer precursors, 3‐(Ar2OCH2)‐2‐Br‐(Ar1N?CH)C6H3 ((Ar1NCOAr2)Br, 3a , 3b , 3c , 3d ) and 3‐(2,6‐Me2C6H3SCH2)‐2‐Br‐(Ar1N?CH)C6H3 ((Ar1NCSMe)Br, 4a and 4b ) were synthesized and characterized. The reactions of [Ar1NCOAr2]Br/ [Ar1NCSMe]Br with nBuLi and the subsequent addition of the rare‐earth‐metal chlorides afforded their corresponding rare‐earth‐metal–pincer complexes, that is, [(Ar1NCOAr2)YCl2(thf)2] ( 5a , 5b , 5c , 5d ), [(Ar1NCOAr2)LuCl2(thf)2] ( 6a , 6d ), [(Ar1NCOAr2)GdCl2(thf)2] ( 7 ), [{(Ar1NCSMe)Y(μ‐Cl)}2{(μ‐Cl)Li(thf)2(μ‐Cl)}2] ( 8 , 9 ), and [{(Ar1NCSMe)Gd(μ‐Cl)}2{(μ‐Cl)Li(thf)2(μ‐Cl)}2] ( 10 , 11 ). These diamagnetic complexes were characterized by 1H and 13C NMR spectroscopy and the molecular structures of compounds 5a , 6a , 7 , and 10 were well‐established by X‐ray diffraction analysis. In compounds 5a , 6a , and 7 , all of the metal centers adopted distorted pentagonal bipyramidal geometries with the NCO donors and two oxygen atoms from the coordinated THF molecules in equatorial positions and the two chlorine atoms in apical positions. Complex 10 is a dimer in which the two equal moieties are linked by two chlorine atoms and two Cl? Li? Cl bridges. In each part, the gadolinium atom adopts a distorted pentagonal bipyramidal geometry. Activated with alkylaluminum and borate, the gadolinium and yttrium complexes showed various activities towards the polymerization of isoprene, thereby affording highly cis‐1,4‐selective polyisoprene, whilst the NCO? lutetium complexes were inert under the same conditions.  相似文献   

12.
To clarify the nature of the Mo?Carene interaction in terphenyl complexes with quadruple Mo?Mo bonds, ether adducts of composition [Mo2(Ar′)(I)(O2CR)2(OEt2)] have been prepared and characterized (Ar′=ArXyl2, R=Me; Ar′=ArMes2, R=Me; Ar′=ArXyl2, R=CF3) (Mes=mesityl; Xyl=2,6‐Me2C6H3, from now on xylyl) and their reactivity toward different neutral Lewis bases investigated. PMe3, P(OMe)3 and PiPr3 were chosen as P‐donors and the reactivity studies complemented with the use of the C‐donors CNXyl and CN2C2Me4 (1,3,4,5‐tetramethylimidazol‐2‐ylidene). New compounds of general formula [Mo2(Ar′)(I)(O2CR)2( L )] were obtained, except for the imidazol‐2‐ylidene ligand that yielded a salt‐like compound of composition [Mo2(ArXyl2)(O2CMe)2(CN2C2Me4)2]I. The Mo?Carene interaction in these complexes has been analyzed with the aid of X‐ray data and computational studies. This interaction compensates the coordinative and electronic unsaturation of one of the Mo atoms in the above complexes, but it seems to be weak in terms of sharing of electron density between the Mo and Carene atoms and appears to have no appreciable effect in the length of the Mo?Mo, Mo?X, and Mo? L bonds present in these molecules.  相似文献   

13.
The reactivity of N‐heterocyclic carbenes (NHCs) and cyclic alkyl amino carbenes (cAACs) with arylboronate esters is reported. The reaction with NHCs leads to the reversible formation of thermally stable Lewis acid/base adducts Ar‐B(OR)2⋅NHC ( Add1 – Add6 ). Addition of cAACMe to the catecholboronate esters 4‐R‐C6H4‐Bcat (R=Me, OMe) also afforded the adducts 4‐R‐C6H4Bcat⋅cAACMe ( Add7 , R=Me and Add8 , R=OMe), which react further at room temperature to give the cAACMe ring‐expanded products RER1 and RER2 . The boronate esters Ar‐B(OR)2 of pinacol, neopentylglycol, and ethyleneglycol react with cAAC at RT via reversible B−C oxidative addition to the carbene carbon atom to afford cAACMe(B{OR}2)(Ar) ( BCA1 – BCA6 ). NMR studies of cAACMe(Bneop)(4‐Me‐C6H4) ( BCA4 ) demonstrate the reversible nature of this oxidative addition process.  相似文献   

14.
A series of α‐keto‐β‐diimine nickel complexes (Ar‐N = C(CH3)‐C(O)‐C(CH3)=N‐Ar)NiBr2; Ar = 2,6‐R‐C6H3‐, R = Me, Et, iPr, and Ar = 2,4,6‐Me3‐C6H3‐) was prepared. All corresponding ligands are unstable even under an inert atmosphere and in a freezer. Stable copper complex intermediates of ligand synthesis and ethyl substituted nickel complex were isolated and characterized by X‐ray. All nickel complexes were used for the polymerization of ethene, propylene, and hex‐1‐ene to investigate their livingness and the extent of chain‐walking. Low‐temperature propene polymerization with less bulky ortho‐substituents was less isospecific than the one with isopropyl derivative. Propene stereoblock copolymers were prepared by iPr derivative combining the polymerization at low temperature to prepare isotactic polypropylene (PP) block and at a higher temperature, supporting chain‐walking, to obtain amorphous regioirregular PP block. Alternatively, a copolymerization of propene with ethene was used for the preparation of amorphous block. © 2017 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2017 , 55, 2440–2449  相似文献   

15.
A series of new hydroxyindanone-imine ligands [PhN=CC2H3(CH3)C6H2(CH3)OH] (HL1) and [ArN=CC2H3(CH3)C6H2(R)OH] (Ar = 2,6-i-Pr(2)C(6)H(3), R = Me (HL2), R = H (HL3), and R = Cl (HL4)) were synthesized and characterized. Reactions of hydroxyindanone-imines with Ni(OAc)(2).4H(2)O result in the formation of the trinuclear hexa(indanone-iminato)tri(nickel(II)) complex Ni(3)[PhN=CC2H3(CH3)C6H2(CH3)O](6) (1) and the mononuclear bis(indanone-iminato)nickel(II) complexes Ni[ArN=CC2H3(CH3)C6H2(R)O](2) (Ar = 2,6-i-Pr(2)C(6)H(3), R = Me (2), R = H (3), and R = Cl (4)). All nickel complexes were characterized by their IR, NMR spectra and elemental analyses. In addition, X-ray structure analyses were performed for complexes 1 and 2. After being activated with methylaluminoxane (MAO), these nickel(II) complexes can be used as catalysts for the polymerization of methyl methacrylate (MMA) to produce syndiotactic-rich PMMA. Catalytic activities and the degree of syndiotacticity of PMMA have been investigated for various reaction conditions.  相似文献   

16.
Interesting varieties of heterobimetallic mixed-ligand complexes [Zr{M(OPri) n }2 (L)] (where M = Al, n = 4, L = OC6H4CH = NCH2CH2O (1); M = Nb, n = 6, L = OC6H4CH = NCH2CH2O (2); M = Al, n = 4, L = OC10H6CH = NCH2CH2O (3); M = Nb, n = 6, L = OC10H6CH = NCH2CH2O (4)), [Zr{Al(OPri)4}2Cl(OAr)] (where Ar = C6H3Me2-2,5 (5); Ar = C6H2Me-4-Bu2-2,6 (6), [Zr{Al(OPri)4}2(OAr)2] (where Ar = C6H3Me2-2,5 (7); Ar = C6H2Me-4-Bu2-2,6 (8), [Zr{Al(OPri)4}3(OAr)] (where Ar = C6H3Me2-2,5 (9); Ar = C6H3Me2-2,6 (10), [ZrAl(OPri)7-n (ON=CMe2) n ] (where n = 4 (11); n = 7 (12), [ZrAl2(OPri)10-n (ON=CMe2) n ] (where n = 4 (13); n = 6 (14); n = 10 (15) and [Zr{Al(OPri)4}2{ON=CMe(R)} n Cl2–n] [where n = 1, R = Me (16); n = 2, R = Me (17); n = 1, R = Et (18); n = 2, R = Et (19)] have been prepared either by the salt elimination method or by alkoxide-ligand exchange. All of these heterobimetallic complexes have been characterized by elemental analyses, molecular weight measurements, and spectroscopic (I.r., 1H-, and 27Al- n.m.r.) studies.  相似文献   

17.
Experimental and theoretical studies on equilibria between iridium hydride alkylidene structures, [(TpMe2)Ir(H){?C(CH2R)ArO }] (TpMe2=hydrotris(3,5‐dimethylpyrazolyl)borate; R=H, Me; Ar=substituted C6H4 group), and their corresponding hydride olefin isomers, [(TpMe2)Ir(H){R(H)C? C(H)OAr}], have been carried out. Compounds of these types are obtained either by reaction of the unsaturated fragment [(TpMe2)Ir(C6H5)2] with o‐C6H4(OH)CH2R, or with the substituted anisoles 2,6‐Me2C6H3OMe, 2,4,6‐Me3C6H2OMe, and 4‐Br‐2,6‐Me2C6H2OMe. The reactions with the substituted anisoles require not only multiple C? H bond activation but also cleavage of the Me? OAr bond and the reversible formation of a C? C bond (as revealed by 13C labeling studies). Equilibria between the two tautomeric structures of these complexes were achieved by prolonged heating at temperatures between 100 and 140 °C, with interconversion of isomeric complexes requiring inversion of the metal configuration, as well as the expected migratory insertion and hydrogen‐elimination reactions. This proposal is supported by a detailed computational exploration of the mechanism at the quantum mechanics (QM) level in the real system. For all compounds investigated, the equilibria favor the alkylidene structure over the olefinic isomer by a factor of between approximately 1 and 25. Calculations demonstrate that the main reason for this preference is the strong Ir–carbene interactions in the carbene isomers, rather than steric destabilization of the olefinic tautomers.  相似文献   

18.
The synthesis, characterization and methyl methacrylate polymerization behaviors of 2‐(N‐arylimino)pyrrolide nickel complexes are described. The nickel complex [NN]2Ni ( 1 , [NN] = [2‐C(H)NAr‐5‐tBu‐C4H2N]?, Ar = 2,6‐iPr2C6H3) was prepared in good yield by the reaction of [NN]Li with trans‐[Ni(Cl)(Ph)(PPh3)2] in THF. Reaction of [NN]Li with NiBr2(DME) yielded the nickel bromide [NN]Ni(Br)[NNH] ( 2 ). Complexes 1 and 2 were characterized by 1H NMR and IR spectroscopy and elemental analysis, and by X‐ray single crystal analysis. Both complexes, upon activation with methylaluminoxane, are highly active for the polymerization of methyl methacrylate to give high molecular weight polymethylmethacrylate with narrow molecular distributions. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

19.
Treatment of the imines [ArN=CH-CH=NAr] and [ArN=CH-2-py] (Ar=2,6-Pr2iC6H3) with AlMe3 in toluene affords the highly crystalline complexes [AlMe2{ArN-CH2-C(Me)=NAr}] (1) and [AlMe2{ArN-CH(Me)-2-py}] (2); the molecular structures of 1 and 2 show that the aluminiums are bonded to imino-amide and pyridyl-amide ligands respectively arising from methyl group transfer from the aluminium centre to the backbone carbon of the imine ligand.  相似文献   

20.
Treatment of pyridine‐stabilized silylene complexes [(η5‐C5Me4R)(CO)2(H)W?SiH(py)(Tsi)] (R=Me, Et; py=pyridine; Tsi=C(SiMe3)3) with an N‐heterocyclic carbene MeIiPr (1,3‐diisopropyl‐4,5‐dimethylimidazol‐2‐ylidene) caused deprotonation to afford anionic silylene complexes [(η5‐C5Me4R)(CO)2W?SiH(Tsi)][HMeIiPr] (R=Me ( 1‐Me ); R=Et ( 1‐Et )). Subsequent oxidation of 1‐Me and 1‐Et with pyridine‐N‐oxide (1 equiv) gave anionic η2‐silaaldehydetungsten complexes [(η5‐C5Me4R)(CO)2W{η2‐O?SiH(Tsi)}][HMeIiPr] (R=Me ( 2‐Me ); R=Et ( 2‐Et )). The formation of an unprecedented W‐Si‐O three‐membered ring was confirmed by X‐ray crystal structure analysis.  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