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 234 毫秒
1.
Single‐site, well‐defined, silica‐supported tantallaaziridine intermediates [≡Si‐O‐Ta(η2‐NRCH2)(NMe2)2] [R=Me ( 2 ), Ph ( 3 )] were prepared from silica‐supported tetrakis(dimethylamido)tantalum [≡Si‐O‐Ta(NMe2)4] ( 1 ) and fully characterized by FTIR spectroscopy, elemental analysis, and 1H,13C HETCOR and DQ TQ solid‐state (SS) NMR spectroscopy. The formation mechanism, by β‐H abstraction, was investigated by SS NMR spectroscopy and supported by DFT calculations. The C?H activation of the dimethylamide ligand is favored for R=Ph. The results from catalytic testing in the hydroaminoalkylation of alkenes were consistent with the N‐alkyl aryl amine substrates being more efficient than N‐dialkyl amines.  相似文献   

2.
The C? C bond forming catalytic hydroaminoalkylation of terminal alkenes, 1,3‐dienes, or styrenes allows a direct and highly atom efficient (100 %) synthesis of amines which can result in the formation of two regioisomers, the linear and the branched product. We present a new titanium catalyst with 2,6‐bis(phenylamino)pyridinato ligands for intermolecular hydroaminoalkylation reactions of styrenes and 1‐phenyl‐1,3‐butadienes that delivers the corresponding linear hydroaminoalkylation products with excellent regioselectivities.  相似文献   

3.
Alkyl‐substituted η5‐pentadienyl half‐sandwich complexes of cobalt have been reported to undergo [5+2] cycloaddition reactions with alkynes to provide η23‐cycloheptadienyl complexes under kinetic control. DFT studies have been used to elucidate the mechanism of the cyclization reaction as well as that of the subsequent isomerization to the final η5‐cycloheptadienyl product. The initial cyclization is a stepwise process of olefin decoordination/alkyne capture, C? C bond formation, olefin arm capture, and a second C? C bond formation; the initial decoordination/capture step is rate‐limiting. Once the η23‐cycloheptadienyl complex has been formed, isomerization to η5‐cycloheptadienyl again involves several steps: olefin decoordination, β‐hydride elimination, reinsertion, and olefin coordination; also here the initial decoordination step is rate limiting. Substituents strongly affect the ease of reaction. Pentadienyl substituents in the 1‐ and 5‐positions assist pentadienyl opening and hence accelerate the reaction, while substituents at the 3‐position have a strongly retarding effect on the same step. Substituents at the alkyne (2‐butyne vs. ethyne) result in much faster isomerization due to easier olefin decoordination. Paths involving triplet states do not appear to be competitive.  相似文献   

4.
The reaction of η51‐pentafulvene titanium complexes with the strong N‐heterocyclic carbene (NHC) donor 1,3,4,5‐tetramethylimidazole‐2‐ylidene, leads to the formation of isolable NHC titanium adducts, featuring a haptotropic shift of the pentafulvene ligand, proved by single crystal X‐ray diffraction as well as NMR spectroscopy studies.  相似文献   

5.
Reactions of pyrimidine‐2‐thione (HpymS) with PdII/PtIV salts in the presence of triphenyl phosphine and bis(diphenylphosphino)alkanes, Ph2P‐(CH2)m‐PPh2 (m = 1, 2) have yielded two types of complexes, viz. a) [M(η2‐N, S‐ pymS)(η1‐S‐ pymS)(PPh3)] (M = Pd, 1 ; Pt, 2 ), and (b) [M(η1‐S‐pymS)2(L‐L)] {L‐L, M = dppm (m = 1) Pd, 3 ; Pt, 4 ; dppe (m = 2), Pd, 5 ; Pt, 6 }. Complexes have been characterized by elemental analysis (C, H, N), NMR spectroscopy (1H, 13C, 31P), and single crystal X‐ray crystallography ( 1 , 2 , 4 , and 5 ). Complexes 1 and 2 have terminal η1‐S and chelating η2‐N, S‐modes of pymS, while other Pd/Pt complexes have only terminal η1‐S modes. The solution state 31P NMR spectral data reveal dynamic equilibrium for the complexes 3 , 5 and 6 , whereas the complexes 1 , 2 and 4 are static in solution state.  相似文献   

6.
Cyclopentadienyl–ruthenium half‐sandwich complexes with η2‐bound alkyne ligands have been suggested as catalytic intermediates in the early stages of Ru‐catalyzed reactions with alkynes. We show that electronically unsaturated complexes of the formula [RuCl(Cp^)(η2‐RC≡CR′)] can be stabilized and crystallized by using the sterically demanding cyclopentadienyl ligand Cp^ (Cp^=η5‐1‐methoxy‐2,4‐tert‐butyl‐3‐neopentyl‐cyclopentadienyl). Furthermore we demonstrate that [RuCl2(Cp^)]2 is an active and regioselective catalyst for the [2+2+2] cyclotrimerization of alkynes. The first elementary steps of the reaction of mono(η2‐alkyne) complexes containing {RuCl(Cp*)} (Cp*=η5‐C5Me5) and {RuCl(Cp^)} fragments with alkynes were investigated by DFT calculations at the M06/6‐31G* level in combination with a continuum solvent model. Theoretical results are able to rationalize and complement the experimental findings. The presence of the sterically demanding Cp^ ligand increases the activation energy required for the formation of the corresponding di(η2‐alkyne) complexes, enhancing the initial regioselectivity, but avoiding the evolution of the system towards the expected cyclotrimerization product when bulky substituents are present. Theoretical results also show that the electronic structure and stability of a metallacyclic intermediate is strongly dependent on the nature of the substituents present in the alkyne.  相似文献   

7.
Metallocene dihalides and derivatives thereof are of great interest as precursors for catalysts in polymerization reactions, as antitumor agents and, due to their increased stability, as suitable starting materials in salt metathesis reactions and the generation of metallocene fragments. We report the synthesis and structural characterization of a series of eleven substituted bis(η5‐cyclopentadienyl)titanium dihalides, namely bis[η5‐1‐(diphenylmethyl)cyclopentadienyl]difluoridotitanium(IV), [Ti(C18H15)2F2], bis{η5‐1‐[bis(4‐methylphenyl)methyl]cyclopentadienyl}difluoridotitanium(IV), [Ti(C20H19)2F2], and bis{η5‐1‐[bis(adamantan‐2‐yl)methyl]cyclopentadienyl}difluoridotitanium(IV), [Ti(C15H19)2F2], together with the bromide and iodide analogues, and the chloride analogues of the diphenylmethyl and adamantyl complexes. These eleven complexes were prepared by the reaction of the corresponding bis(η51‐pentafulvene)titanium complexes with different hydrogen halides (Cl, Br and I). The titanocene fluorides become available via chloride–fluoride exchange reactions.  相似文献   

8.
The synthesis and crystal structures of two dinuclear titanocene hydride complexes are reported. Both complexes, namely bis(η5‐(di‐para‐tolylmethyl)cyclopentadienyl)titanium hydride dimer, [(η5‐C20H19)2Ti(μ‐H)]2 ( 2a ), and bis(η5‐2‐adamantylcyclopentadienyl)‐titanium hydride dimer, [(η5‐C15H19)2Ti(μ‐H)]2 ( 2b ), are formed via activation of molecular hydrogen by the corresponding bis(η51‐pentafulvene)titanium complexes 1a and 1b at ambient temperatures and pressures in high yields. The hydride complexes 2a and 2b exhibit planar [Ti2H2] cores and, as a result of the heterolytic cleavage of molecular hydrogen, substituted Cp Ligands were formed during the reaction.  相似文献   

9.
Transition‐metal carbene complexes and their reactivities are a key topic of chemistry. They are an integral part of researches in catalysis, organic synthesis, coordination chemistry, and numerous other areas. In this context, we report the synthesis of a low‐valent bis(η51‐(di‐p‐tolyl)‐pentafulvene)niobium chloride. Owing to the π‐η5:σ‐η1 coordination mode of the pentafulvenes and the resulting high nucleophilic character of the exocyclic carbon atom of the ligand, the bis(η51‐pentafulvene)niobium complex is able to achieve the umpolung of a coordinated vinyl unit and the resulting formation of the first η51 cyclic niobium Schrock carbene complex. This new synthetic route is, in comparison to classical α‐hydrogen elimination reactions or thermolysis of diazo compounds, completely unprecedented. The reactivity of the cyclic carbene function and the remaining fulvene ligand is demonstrated by double N?H bond activation of primary amines to niobium imido complexes.  相似文献   

10.
Syntheses and Structures of η1‐Phosphaallyl, η1‐Arsaallyl, and η1‐Stibaallyl Iron Complexes [(η5‐C5Me5)(CO)2Fe–E(SiMe3)C(OSiMe3)=CPh2] (E = P, As, Sb) The reaction of equimolar amounts of [(η5‐C5Me5)(CO)2Fe–E(SiMe3)2] ( 1 a : E = P; 1 b : As; 1 c : Sb) and diphenylketene afforded the η1‐phosphaallyl‐, η1‐arsaallyl‐, and η1‐stibaallyl complexes [(η5‐C5Me5)(CO)2Fe–E(SiMe3)C(OSiMe3)=CPh2] ( 2 a : E = P; 2 b : As; 2 c : Sb). The molecular structures of 2 b and 2 c were elucidated by single crystal X‐ray analyses.  相似文献   

11.
Two ligand exchange reactions at the titanium atom in quasi-tetrahedral titanocene complexes have been studied. The first is substitution of a Cl ligand by an aryloxy group starting from substrates η5-Cp-η5-Cp′Ti(Cl)OPh which have an planar chirality on the Cp′ ring. The second is the substitution of one of the aryloxy groups of the complexes η5-Cp-η5-Cp′Ti(OPh′)OPh by the action of HCl. In this case, the reaction is generally selective and has a high degree of stereo-specificity with retention at the titanium atom. This retention has been established by crystallographic analysis of two suitable substrates: diastereoisomer F. 171°C of η5-C5H55-(1-Me-3-CHMe2C5H5)](2-ClC6H4O)(2,6-Me2C6H3O)Ti and diastereoisomer F. 134°C of η5-C5H55-(1-Me-3-CHMe2C5H3)](2-ClC6H4O)TiCl.  相似文献   

12.
Recently, esters have received much attention as transmetalation partners for cross‐coupling reactions. Herein, we report a systematic study of the reactivity of a series of esters and thioesters with [{(dtbpe)Ni}2(μ‐η22‐C6H6)] (dtbpe=1,2‐bis(di‐tert‐butyl)phosphinoethane), which is a source of (dtbpe)nickel(0). Trifluoromethylthioesters were found to form η2‐carbonyl complexes. In contrast, acetylthioesters underwent rapid Cacyl?S bond cleavage followed by decarbonylation to generate methylnickel complexes. This decarbonylation could be pushed backwards by the addition of CO, allowing for regeneration of the thioester. Most of the thioester complexes were found to undergo stoichiometric cross‐coupling with phenylboronic acid to yield sulfides. While ethyl trifluoroacetate was also found to form an η2‐carbonyl complex, phenyl esters were found to predominantly undergo Caryl?O bond cleavage to yield arylnickel complexes. These could also undergo transmetalation to yield biaryls. Attempts to render the reactions catalytic were hindered by ligand scrambling to yield nickel bis(acetate) complexes, the formation of which was supported by independent syntheses. Finally, 2‐naphthyl acetate was also found to undergo clean Caryl?O bond cleavage, and although stoichiometric cross‐coupling with phenylboronic acid proceeded with good yield, catalytic turnover has so far proven elusive.  相似文献   

13.
Two unsymmetrical 1,2,4‐trithiolanes ( 1d and 1e ) were reacted with [Pt0(PPh3)22‐nbe)] ( 6 ; nbe=norborn‐2‐ene) and [Pt0(dppn)(η2‐nbe)] ( 11 ; dppn=1,8‐bis(diphenylphosphanyl)naphthalene)), respectively. Their treatment with compound 6 resulted in the formation of six‐membered platinacycles of type 7 , which selectively underwent fragmentation into dithiolato complexes and thiobenzophenone ( 4b ). The isolation of the first stable C‐substituted member of this class of compounds (i.e., compound 7e ) permitted kinetic studies of this process by using UV/Vis spectroscopy. These results, together with DFT calculations, allowed us to propose a mechanism for the reactions of compound 6 with 1,2,4‐trithiolanes. In contrast, similar treatment of compound 11 with compounds 1d and 1e at room temperature did not result in any reaction. Heating the appropriate samples to 110 °C led to the formation of dithiolato complexes and η2‐thioketone compounds, thus pointing to a thermally induced [3+2]‐cycloreversion of the heterocycles as an initial step of the reaction.  相似文献   

14.
Cobalt‐catalyzed cross‐dimerization of simple alkenes with 1,3‐enynes is reported. A [2+2] cycloaddition reaction occurred, with alkenes bearing no allylic hydrogen, by reductive elimination of a η3‐butadienyl cobaltacycle. On the other hand, aliphatic alkenes underwent 1,4‐hydroallylation by means of exo‐cyclic β‐H elimination. These reactions can provide cyclobutenes and allenes that were previously difficult to access, from simple substrates in a highly chemo‐ and regioselective manner.  相似文献   

15.
The syndiospecific polymerization of styrene was investigated with the fluorine‐containing half‐sandwich complexes η5‐pentamethylcyclopentadienyl titanium bis(trifluoroacetate) dimer, η5‐octahydrofluorenyl titanium tristrifluoro‐acetate, η5‐octahydrofluorenyl titanium dimethoxymonotrifluoroacetate, and η5‐octahydrofluorenyl titanium tris(pentafluorobenzoate) in comparison to known chloride and methoxide complexes in the presence of relatively low amounts of methylalumoxane and triisobutylaluminum. After the selection of effective reaction conditions for a solvent‐free polymerization, the following orders of decreasing polymerization activity of the titanium complexes can be observed: for pentamethylcyclopentadienyl compounds, Cp*Ti(OMe)3 > [Cp*Ti(OCOCF3)2]2O ≈ Cp*TiCl3, and for octahydrofluorenyl compounds, [656]Ti(OMe)3 > [656]Ti(OCOC6F5)3 > [656]Ti(OCH3)2(OCOCF3) > [656]Ti (OCOCF3)3. The [656]Ti complexes, showing the highest polymerization conversions at 70 °C and in comparison with the Cp* Ti compounds, turned out to be highly efficient catalysts for the syndiospecific styrene polymerization. The fluorine‐containing Cp* and [656]Ti complexes lead to much higher molecular weights than the chloride and methoxide compounds because of a reduction in chain‐limiting transfer reactions. The introduction of only one fluorine‐containing ligand into the coordination sphere of the metal compound is obviously sufficient for a significant increase in molecular weight. The active polymerization sites of the [656]Ti complexes with methylalumoxane and triisobutylaluminum are extremely stable during storage at room temperature in regard to their polymerization activity. © 2000 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 38: 2428–2439, 2000  相似文献   

16.
The η1-diacetylenic molybdenum complexes η-C5H5(CO)3MoCH2CCCCCH3 (Ia) can add two methanoi molecules successively. The first addition, with CO insertion, gives a usual η3-allyl-alkoxycarbonylated compound, gh5-C5H5(CO)2Mo-η3CH2C(COOCH3)CHCCCH3 (IIa). The second reaction needs propargyl bromide as catalyst. It is a 1,5-methanol addition on the unsaturated η3-allyl ligand to give the new complex η5-C5H5(CO)2Mo-η3-CH3OCH2C(COOCH3)CHCCHCH3 (IIIa). The synthesis of the two unstable cationic intermediates and their reactions with methoxide yielding the same addition products have been achieved and have confirmed the mechanism postulated. With the iron analogue (η5-C5H5)(CO)2FeCH2CCCCCH3 (Ib), direct addition of methanoi is not possible, but the same reactions are obtained by protonation followed by methoxide addition to give complexes IIb and IIIb. In this case, the more stable cationic intermediates IVb and Vb can be fully characterised.  相似文献   

17.
Cyclopropanecarboxaldehyde ( 1 a ), cyclopropyl methyl ketone ( 1 b ), and cyclopropyl phenyl ketone ( 1 c ) were reacted with [Ni(cod)2] (cod=1,5‐cyclooctadiene) and PBu3 at 100 °C to give η2‐enonenickel complexes ( 2 a – c ). In the presence of PCy3 (Cy=cyclohexyl), 1 a and 1 b reacted with [Ni(cod)2] to give the corresponding μ‐η21‐enonenickel complexes ( 3 a , 3 b ). However, the reaction of 1 c under the same reaction conditions gave a mixture of 3 c and cyclopentane derivatives ( 4 c , 4 c′ ), that is, a [3+2] cycloaddition product of 1 c with (E)‐1‐phenylbut‐2‐en‐1‐one, an isomer of 1 c . In the presence of a catalytic amount of [Ni(cod)2] and PCy3, [3+2] homo‐cycloaddition proceeded to give a mixture of 4 c (76 %) and 4 c′ (17 %). At room temperature, a possible intermediate, 6 c , was observed and isolated by reprecipitation at ?20 °C. In the presence of 1,3‐bis(2,6‐diisopropylphenyl)imidazol‐2‐ylidene (IPr), both 1 a and 1 c rapidly underwent oxidative addition to nickel(0) to give the corresponding six‐membered oxa‐nickelacycles ( 6 ai , 6 ci ). On the other hand, 1 b reacted with nickel(0) to give the corresponding μ‐η21‐enonenickel complex ( 3 bi ). The molecular structures of 6 ai and 6 ci were confirmed by X‐ray crystallography. The molecular structure of 6 ai shows a dimeric η1‐nickelenolate structure. However, the molecular structure of 6 ci shows a monomeric η1‐nickelenolate structure, and the nickel(II) 14‐electron center is regarded as having “an unusual T‐shaped planar” coordination geometry. The insertion of enones into monomeric η1‐nickelenolate complexes 6 c and 6 ci occurred at room temperature to generate η3‐oxa‐allylnickel complexes ( 8 , 9 ), whereas insertion into dimeric η1‐nickelenolate complex 6 ai did not take place. The diastereoselectivity of the insertion of an enone into 6 c having PCy3 as a ligand differs from that into 6 ci having IPr as a ligand. In addition, the stereochemistry of η3‐oxa‐allylnickel complexes having IPr as a ligand is retained during reductive elimination to yield the corresponding [3+2] cycloaddition product, which is consistent with the diastereoselectivity observed in Ni0/IPr‐catalyzed [3+2] cycloaddition reactions of cyclopropyl ketones with enones. In contrast, reductive elimination from the η3‐oxa‐allylnickel having PCy3 as a ligand proceeds with inversion of stereochemistry. This is probably due to rapid isomerization between syn and anti isomers prior to reductive elimination.  相似文献   

18.
The mechanism of the carbonylation reaction of allyl halides catalyzed by nickel (Ni(CO)4) and palladium (Cl2Pd(PPh3)2) complexes is theoretically investigated at the DFT level using the hybrid B3LYP functional. The favored reaction path to carbonylation corresponds, for both catalysts, to a direct attack of the halogen on the metal. This affords η1 intermediates that can undergo the final carbonylation step. It is also possible to obtain the acyl product (β,γ-unsaturated acyl halides) from η2 and/or η3 intermediates. However, in this case, the barrier of the rate-determining step to carbonylation is much higher. Since a channel on the potential surface connects rather easily the η2 or η3 intermediates to the η1 intermediates, an alternative and competitive path leading to the acyl products can originate from the η2 or η3 intermediates, follow the η23 → η1 transformation, then undergo the final carbonylation step.  相似文献   

19.
Addition of PR3 (R=Ph or OPh) to [Cu(η2‐Me6C6)2][PF6] results in the formation of [(η6‐Me6C6)Cu(PR3)][PF6], the first copper–arene complexes to feature an unsupported η6 arene interaction. A DFT analysis reveals that the preference for the η6 binding mode is enforced by the steric clash between the methyl groups of the arene ligand and the phenyl rings of the phosphine co‐ligand.  相似文献   

20.
The first two‐dimensional lanthanum(III) coordination polymer, [La(1,5‐NDS)1.5(H2O)5]n (1) (1,5‐NDS2? = 1,5‐naphthalenedisulfonate), was synthesized and structurally characterized by single‐crystal X‐ray diffraction analysis. The disulfonate ligands act in the η112 and η1113 binding modes to link the LaO9 tricapped trigonal prisms into a lamellar structure. It exhibits strong purple emission in the solid state.  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