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 reaction of a series of β-methoxyvinyl trifluoromethyl ketones [CF3COC(R2)?C(OMe)R1, where R1 = Me, -(CH2)3-C3, -CH2)4-C3, Ph and R2 = H, Me, -(CH2)3-C4, -(CH2)4-C4] with N-methylhydroxylamine is reported. The regiochemistry of the reaction are explained by MO calculation data.  相似文献   

2.
η2-Acyl and σ-Alkyl(carbonyl) Coordination in Molybdenum and Tungsten Complexes: Synthesis and Studies of the Isomerization Equilibria and Kinetics The anionic molybdenum and tungsten complexes [LRM(CO)3]? (LR? = [(C5H5)Co{P(O)R2}3]?, R = OCH3, OC2H5, O-i-C3H7; M = Mo, W) have been alkylated with the iodides R′ I, R′ = CH3, C2H5, i-C3H7, and CH2C6H5. The reactivity pattern of the alkylation is in accord with a SN2 mechanism. Depending on M, R′, reaction temperature, and time the η-alkyl (carbonyl) compounds [LRM(CO)3R′] and/or the isomeric η2-acyl compounds [LRM(CO)22-COR′)] can be obtained. 8 new σ-alkyl(carbonyl) compounds and 15 new η2-acyl compounds have been isolated and characterized. The 1H NMR and the IR spectra give conclusive evidence that the σ-alkyl(carbonyl) compounds [LRM(CO)3R′] are formed as the primary products of the alkylation and that they isomerize partly or completely to give the η2-acyl compounds [LRM(CO)22-COR′)]. The position of the equilibrium σ-alkyl(carbonyl)/η2-acyl is controlled by the steric demands of the groups R′ and the ligands LR?. The molybdenum compounds isomerize much more readily than the tungsten compounds. The rate constants of the isomerization processes [LRMo(CO)3CH3] → [LRMo(CO)22-COCH3)], R = OCH3, OC2H5, and O-i-C3H7, measured at 305 K in acetone-d6, are 6–8 x 10?3 s?1.  相似文献   

3.
Reactions of CH3OCH2CH2OH, PhOCH2CH2OH, o-CH3OC6H4CH2OH, or PhSCH2CH2OH with Cp2ZrCl2 in the presence of NEt3 gave series monomeric complexes Cp2ZrClX (X = OCH2CH2OCH3 ( 1 ), OCH2CH2OPh ( 2 ), o-OCH2C6H4OCH3 ( 3 ) or OCH2CH2SPh ( 4 )). In a reaction of Cp2TiCl2 with PhSCH2CH2OH in the presence of NEt3, the complex Cp2TiCl(OCH2CH2SPh) ( 5 ) was obtained. These complexes were characterized by the 1H and 13C NMR, and the chemical shifts of the Cp rings for complexes 1–4 are nearly identical, despite differences in ligands for an indication of similar structures of these complexes around the metal center. Complex 4 crystallizes into the monoclinic P21/c space group with a = 13.806(3) Å, b = 16.394(3) Å, c = 7.838(2) Å, β = 100.85(3)° Z = 4, R = 0.029, Rw = 0.044, and Gof = 1.19. The molecular structure of complex 4 shows that the 2-(phenylthio)ethoxide bonds to the Zr metal center through the alkoxide donor leaving the thioether donor free from coordination.  相似文献   

4.
N‐(Dialkylthiocarbamoyl)benzimidoyl chlorides react with o‐(salicylidenimine)benzylamine with formation of a novel class of tetradentate benzamidine ligands (H2LEt and H2LMorph), which readily react with Ni(CH3COO)2, [PdCl2(CH3CN)2], and [PtCl2(PPh3)2] under formation of complexes of the composition [M(LR)] [M = Ni ( 4 ), Pd ( 5 ), Pt ( 6 )]. In all complexes, H2LR is doubly deprotonated and bonded to the metal ion via its N2OS donor set and establishes a distorted square‐planar coordination sphere. The antiproliferative effects of the compounds on MCF‐7 and Hep‐G2 cells were studied. The complexes of H2LMorph are generally more active than those of H2LEt. While H2LEt and its complexes exhibit stronger effects on the Hep‐G2 line, the corresponding compounds of H2LMorph show almost equal effects on the two cell lines. In each series of compounds, the cytotoxicity increases in the order H2LR << 4 < 5 < 6 .  相似文献   

5.
New dinuclear pentacoordinate molybdenum(V) complexes, [Mo2VO3L2] [L = thiosemicarbazonato ligand: C6H4(O)CH:NN:C(S)NHR′ and C10H6(O)CH:NN:C(S)NHR′; R′ = H, CH3, C6H5) were obtained either by oxygen atom abstraction from MoVIO2L with triphenylphosphine or by using [Mo2O3(acac)4] in the reaction with the corresponding ligands H2L. Crystal and molecular structure of [Mo2O3{C6H4(O)CH:NN:C(S)NHC6H5}2] · CH3CN has been determined by the single‐crystal X‐ray diffraction method.  相似文献   

6.
The synthesis of new functionalized organotin‐chalcogenide complexes was achieved by systematic optimization of the reaction conditions. The structures of compounds [(R1, 2Sn)3S4Cl] ( 1 , 2 ), [((R2Sn)2SnS4)2(μ‐S)2] ( 3 ), [(R1, 2Sn)3Se4][SnCl3] ( 4, 5 ), and [Li(thf)n][(R3Sn)(HR3Sn)2Se4Cl] ( 6 ), in which R1=CMe2CH2C(O)Me, R2=CMe2CH2C(NNH2)Me, and R3=CH2CH2COO, are based on defect heterocubane scaffolds, as shown by X‐ray diffraction, 119Sn NMR spectroscopy, and ESI mass spectrometry analyses. Compounds 4 , 5 , and 6 constitute the first examples of defect heterocubane‐type metal‐chalcogenide complexes that are comprised of selenide ligands. Comprehensive DFT calculations prompted us to search for the formal intermediates [(R1SnCl2)2(μ‐S)] ( 7 ) and [(R1SnCl)2(μ‐S)2] ( 8 ), which were isolated and helped to understand the stepwise formation of compounds 1 – 6 .  相似文献   

7.
The treatment of various allylic chlorides or bromides with zinc dust in the presence of lithium chloride and magnesium pivalate (Mg(OCOtBu)2) in THF affords allylic zinc reagents which, after evaporation of the solvent, produce solid zinc reagents that display excellent thermal stability. These allylic reagents undergo Pd‐catalyzed cross‐coupling reactions with PEPPSI‐IPent, as well as highly regioselective and diastereoselective additions to aryl ketones and aldehydes. Acylation with various acid chlorides regioselectively produces the corresponding homoallylic ketones, with the new C? C bond always being formed on the most hindered carbon of the allylic system.  相似文献   

8.
Hydrazido(2?) and hydrazido(1?) complexes of tungsten condense with ketones, R1R2CO, in the presence of catalytic amounts of acid to yield complexes containing the groups W = NN = CR1R2 and WNHN =CR1R2 respectively. The other ligands are halide ions and monotertiary phosphines. These new complexes yield secondary amines and ammonia on reduction with LiAlH4; acids produce nitrogen-free tungsten materials, hydrazine and azines.  相似文献   

9.
Pyridine N-imine complexes of methylcobaloxime, CH3Co(Hdmg)2(R1— C5HnN+N?H) (n = 4; R1 = H, 2-CH3, 3-CH3, 4-CH3: n = 3; R1 = 2,6-CH3), have been synthesized by the reaction of CH3Co(Hdmg)2S(CH3)2 with a pyridine N-imine which is generated from a pyridine, hydroxylamine-O-sulfonic acid and K2CO3. The reactions of CH3Co(Hdmg)2(C5H5N+N?H) with acid anhydrides form new methylcobaloxime complexes with N-substituted pyridine N-imines, CH3Co(Hdmg)2(C5H5N+N?R2) R2 = COPh, COMe, COEt). With maleic anhydride, (pyridine N-acryloylimine)carboxylic acid is formed. With acetylenedicarboxylic acid dimethyl ester, 1,3-dipolar cycloaddition of the ligand gives pyrazolo[1,5-a]pyridine-2,3-dicarboxylic acid dimethyl ester.  相似文献   

10.
The complexes formed by combining Pd(OAc)2 and iminophosphine ligands (P^N) are active catalysts in Suzuki–Miyaura cross-coupling reactions under mild conditions. Aryl bromides and iodides, as well as benzyl chlorides give the corresponding coupled products in high yields at low temperatures (25–50 °C) using these catalysts. Iminophosphines containing the most sterically demanding groups attached to the N-imino moiety were the most effective ligands. New divalent Pd complexes of known iminophosphines were synthesised and their activity was compared with the in situ generated catalyst system. The complex resulting from the oxidative addition of 4-bromo anisole [Pd(4-CH3OC6H4)Br(P^N)] was more active than the in situ generated system. However, palladacycles containing the iminophosphine ligand (e.g., {[C6H4CH(Me)2St-Bu]Pd(P^N)}+PF6) were less active than the in situ generated catalyst due to the greater stability of the complexes that involve two bidentate ligands. Poisoning tests demonstrated that homogeneous mononuclear palladium species containing the iminophsophine ligand were responsible for the catalytic activity.  相似文献   

11.
Reactions of N-(2-hydroxy-3,5-R1,R2-benzyl)-4-aminoantipyrines with copper acetate in ethanol gave complexes with Schiff bases (SBs) rather than the expected complexes with reduced SBs; i.e., the starting ligands undergo oxidative dehydrogenation during the complexation reaction. The corresponding complexes with reduced SBs were obtained from sodium salts of the ligands and cupric sulfate in aqueous solutions. Kinetic measurements showed that oxidative dehydrogenation occurs in the heteroleptic complexes Cu(L i )(CH3COO)(X) (L i H are derivatives of N-(2-hydroxy-3,5-R1,R2-benzyl)-4-aminoantipyrines; i = 6–10; X = H2O, CH3OH, CH3CH2OH) but does not occur in the complexes CH3OH, CH3CH2OH. The absence of oxidative dehydrogenation of the ligands in Cu(L i )2 · H2O can be explained by the octahedral environment of the Cu2+ ion and, accordingly, the absence of the coordination site for molecular oxygen. The molecular structures of two Cu(II) complexes with SBs were determined by X-ray diffraction.  相似文献   

12.
Kazuhide Kataoka 《Tetrahedron》2006,62(11):2471-2483
Prins cyclization reaction (PCR) of optically active homoallylic alcohols, RaC*H(OH)CH2CHCHCH3 (1-substituted but-2-en-1-ol), with aldehydes (RbCHO) in the presence of an acid-catalyst (HX) affords (2-Rb,3-CH3,4-X,6-Ra)-tetrasubstituted tetrahydropyrans highly stereoselectively in good yields.  相似文献   

13.
Investigations on the Coordination Chemistry of Zinc Dialkyls. XIV. On Lithium and Zinc 3(N, N-Dialkylamino)propyl Compounds It is reported on synthesis and properties of organo lithium compounds of the type [R2NCH2CH2CH2Li]n. The structure is proposed by reason of molecular weight determination and 13C-NMR spectra. In dependence of the molar ratio the lithium dialkylamino propyls and the corresponding Grignard reagents react with zinc chloride forming dimer alkyl zinc chlorides [R2NCH2CH2CH2ZnCl]2 or monomer spiranoide chelate complexes of the formula [R2NCH2CH2CH2]2Zn (R = CH3, C2H5).  相似文献   

14.
The reaction of CrO2Cl2 with 2, 2′‐bipyridyl or 1, 10‐phenanthroline (diimine) in CCl4 or anhydrous CH3CO2H solution, produces orange‐brown diamagnetic [CrO2Cl2(diimine)]. The X‐ray structure of [CrO2Cl2(2, 2′‐bipy)] shows a six‐coordinate central chromium(VI) atom with cis‐dioxo groups trans to the diimine. In contrast, the diimines react with CrO3 in CH3CO2H / conc. aqueous HCl to form bright red paramagnetic CrV complexes, [CrOCl3(diimine)]. The X‐ray structure of [CrOCl3(2, 2′‐bipy)] shows a six‐coordinate central chromium atom with mer‐chlorines and the diimine trans to O/Cl. The addition of [2, 2‐bipyH2]Cl2 to a solution of CrO3 in CH3CO2H saturated with HCl gas, produces the CrV anion [2, 2′‐bipyH2][CrOCl4]Cl, which loses HCl on heating in vacuo to form [CrOCl3(2, 2′‐bipy)]. IR, UV/Vis, and 1H NMR spectra (CrVI only) are reported for the new complexes. Attempts to extend these routes to oxygen donor ligands, including ethers and phosphine oxides, were unsuccessful. The diimine complexes are the first structurally autheticated adducts of chromium(VI) and (V) oxide‐chlorides with neutral ligands.  相似文献   

15.
Reaction of coordinated (diphenylphosphino)methane and ketones or aldehydes have been characterized by 31P{H1}-NMR, 1H{31P}-NMR, and UV/vis spectroscopy in dichloromethane. Group VI metals hexacarbonyl [M(CO)6 where M = Cr, Mo, and W] reacted with (diphenylphosphino)methane, [(Ph2P)2CH2], to give [(OC)4M{(Ph2P)2CH2}] depending upon the reaction conditions. Condensation of [(CO)4M{(Ph2P)2CH2}] with different ketones or aldehydes forms [(CO)4M{(Ph2P)2C = CR1R2}]. Complexes of the types [(OC)4M{(Ph2P)2C = CR1R2}] reacted with hydrazine in a Michael addition to give [(CO)4M{(Ph2P)2CHC(R1R2)NHNH2}](1.3a–e), which condensed with different ketones and aldehydes to give complex of the type [(CO)4M{(Ph2P)2CHC(R1R2)NHN = C(R3)] (1.4a–e). The structures of the complexes are discussed on the basis of elemental analysis (EA), IR,1H-NMR, 31P-NMR spectroscopic data, and FAB mass spectra. The UV/vis spectra show two absorption bands with the low energy band moving to lower energy with increasing substitution on the (diphenylphosphino) methane (dppm) (a bathochromic effect).  相似文献   

16.
The dipyrromethene (DPM) ligand is the key to isolation of monomeric Zn hydride complexes with tricoordinate zinc centers. A range of RDPM ligands with various substituents in the pole position (1,9-positions) were prepared: R = tBu, adamantyl (Ad), mesityl (Mes), 2,6-diisopropylphenyl (DIPP), 2,4,6-triphenylphenyl (Mes*), or 9-anthracenyl (Anth). Reaction of the ligands with Et2Zn gave a series of (RDPM)ZnEt complexes, which were converted with I2 to the corresponding (RDPM)ZnI compounds. The latter reacted by salt metathesis with KN(iPr)HBH3 to the series of Zn hydride complexes (RDPM)ZnH. For ligands with the larger Mes* and Anth substituents, (RDPM)ZnEt was converted to (RDPM)ZnOSiPh3, which after reaction with PhSiH3 gave the hydrides. While Zn hydride complexes with R = tBu or Ad are dimeric, all complexes with aryl-substituents are monomeric. The aryl groups span a cavity around the metal, blocking dimerization and causing a high-field shift of the 1H NMR signals due to the ASIS effect. Attempted abstraction of the hydride with B(C6F5)3 led to cleavage of the B-C6F5 bond.  相似文献   

17.
《Polyhedron》1986,5(10):1647-1650
[Ni(PnAO)-6H]o, a Ni,2N,3C-pseudoaromatic heterocycle is shown to be highly reactive at the central carbon atom toward most aldehydes and ketones, giving with CH2O, for example, a dimer having the formula [Ni(PnAO)-7H]2-CH2 whose X-ray crystal structure as well as 13C and 1H NMR spectra are presented.  相似文献   

18.
A new family of cationic rhenium tricarbonyl complexes with either two N‐alkylimidazole (N‐RIm) and one pyridine (Py) ligand, or two pyridine and one N‐RIm ligand, [Re(CO)3(N‐RIm)(3?x)(Py)x]+, has been prepared. The reaction of these complexes with a strong base, followed by an oxidant, selectively afforded 2,2’‐pyridylimidazole complexes as the result of intramolecular dehydrogenative C?C coupling reactions. For tris(pyridine) complexes [Re(CO)3(Py)3]+ the reaction pattern upon a deprotonation/oxidation sequence is maintained, which allows the generation of complexes with 2,2’‐bipyridine ligands. In the particular combination of two different types of pyridine ligand in the cationic fac‐Re(CO)3 complexes only the cross‐coupling products with asymmetric 2,2’‐bipyridine ligands were obtained; the homocoupling products were not observed.  相似文献   

19.
 This article gives an overview of recent chemistry based on the tris-acetonitrile complex [RuCp(CH3CN)3]+. Due to the labile nature of the CH3CN ligands, substitution reactions are a dominant feature of this complex. Important derivatives are the highly reactive complexes [RuCp(PR 3)(CH3CN)2]+ which are a source of the 14e fragment [RuCp(PR 3)]+. These species are catalytically active in the redox isomerization of allyl alcohols to give aldehydes and ketones. Furthermore, the cationic complex [RuCp1(P),η2-PPh2CH2CH2CH*CH2)(CH3CN)]PF6 derived from the reaction of [RuCp(CH3CN)3]+ with PPh2CH2CH2CH*CH2 is a model compound for studying coupling reactions of olefins and acetylenes. In addition, [RuCp(CH3CN)3]+ is a valuable precursor for the synthesis of configurationally stable chiral three-legged piano-stool ruthenium complexes. These are currently being intensively investigated as Lewis acid catalysts in asymmetric synthesis.  相似文献   

20.
A series of novel zirconium complexes {R2Cp[2‐R1‐6‐(2‐CH3OC6H4N?CH)C6H3O]ZrCl2 ( 1 , R1 = H, R2 = H, 2 : R1 = CH3, R2 = H; 3 , R1 = tBu, R2 = H; 4 , R1 = H, R2 = CH3; 5 , R1 = H, R2 = n‐Bu)} bearing mono‐Cp and tridentate Schiff base [ONO] ligands are prepared by the reaction of corresponding lithium salt of Schiff base ligands with R2CpZrCl3·DME. All complexes were well characterized by 1H NMR, MS, IR and elemental analysis. The molecular structure of complex 1 was further confirmed by X‐ray diffraction study, where the bond angle of Cl? Zr? Cl is extremely wide [151.71(3)°]. A nine‐membered zirconoxacycle complex Cp(O? 2? C6H4N?CHC6H4‐2? O)ZrCl2 ( 6 ) can be obtained by an intramolecular elimination of CH3Cl from complex 1 or by the reaction of CpZrCl3·DME with dilithium salt of ligand. When activated by excess methylaluminoxane (MAO), complexes 1–6 exhibit high catalytic activities for ethylene polymerization. The influence of polymerization temperature on the activities of ethylene polymerization is investigated, and these complexes show high thermal stability. Complex 6 is also active for the copolymerization of ethylene and 1‐hexene with low 1‐hexene incorporation ability (1.10%). Copyright © 2006 John Wiley & Sons, Ltd.  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