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 46 毫秒
1.
Ruthenium(II) complexes of the type [Ru(PPh3)( 5-C5Me5)L] have been synthesized by the reactions of [RuCl(PPh3)2( 5-C5Me5)] with Schiff bases having the (N, O) donor atoms. The Schiff bases used in this study were prepared by condensing the appropriate aniline with salicylaldehyde or 2-hydroxy-1-naphthaldehyde in a 1:2 molar ratio respectively. The complexes were characterized by analytical, spectral (i.r., electronic and 1H-n.m.r.) data. The new complexes have been used as catalysts in aryl–aryl coupling reactions.  相似文献   

2.
The Zn-Ta, Zn-Nb, Zn-Ti and Zn-Pb systems have been investigated with regard to the formation of mixed-metal alkoxides using infra-red as well as NMR (1H, 207Pb). Lewis acid-base reactions between alkoxides, and between alkoxides and zinc acetate, as well as metathesis reactions, have been considered. Zinc alkoxides are polymeric and quite inert with respect to the formation of heterometallic alkoxides. Their intertness can be overcome in the case of tantalum isopropoxide by using a in situ, freshly prepared colloidal suspension of Zn(OiPr)2. Metathesis reactions are another means of ensuring the formation of mixed-metal zinc-transition metal (M = Ta, Nb, Ti) alkoxides. The solid-state structure determination of [Ta2Zn(-O)(3-O)(-OiPr)3(OiPr)4I]2 2 makes it possible to establish that of the non-halide derivative [ZnTa2O2(OiPr)8]2 1. 1HNMR data indicate that the structures of both aggregates are retained in solution. Hydrolysis-polycondensation reactions of 2 and 1 give after thermal treatment Ta2ZnO6 with traces of Ta2O5 and of Ta2Zn3O8 and pure Ta2ZnO6 respectively.  相似文献   

3.
Anwesha Pal 《Tetrahedron》2010,66(29):5451-5458
Combination of 3,5-dimethyl-1-(2-(diphenylphosphino)phenyl)-1H-pyrazole, A, and 3-tert-butyl-5-methyl-1-(2-(diphenylphosphino)phenyl)-1H-pyrazole, B, with Pd2(dba)3 furnished excellent catalysts for Stille, Kumada and Hiyama cross-coupling reactions. Effects of solvents, bases and ligand/palladium ratio on efficiency of coupling reactions were studied. Molecular structures of palladium(II) complexes of A and B determined by single-crystal X-ray diffraction method revealed a close similarity of ligand arrangement around the metal centre.  相似文献   

4.
The author's work on the reactions of Cp2Mo2Co2S3(CO)4 (1) and Cp2Mo2 Co2S4(CO)2 (2) with organosulfur compounds is reviewed. Reaction pathways that lead to C S bond scission are proposed, and the relexance of these results to HDS reactions mer commercial CoMoS catalysts is discussed.  相似文献   

5.
We extend our sol-gel methodology of one-pot sequences of reactions with opposing reagents to an enzyme/metal-complex pair. Sol-gel entrapped lipase and sol-gel entrapped RhCl[P(C(6)H(5))(3)](3) or Rh(2)Co(2)(CO)(12) were used for one-pot esterification and C-C double bond hydrogenation reactions, leading to saturated esters in good yields. When only the enzyme is entrapped, the homogeneous catalysts quench its activity and poison it. Thus, when 10-undecenoic acid and 1-pentanol were subjected in one pot to the entrapped lipase and to homogeneously dissolved RhCl[P(C(6)H(5))(3)](3) under hydrogen pressure, only 7% of the saturated 1-pentyl undecanoate was obtained. The yield jumped 6.5-fold when both the enzyme and the catalyst were immobilized separately in silica sol-gel matrixes. Similar one-pot esterifications and hydrogenations by sol-gel entrapped lipase and heterogenized rhodium complexes were carried out successfully with the saturated nonoic, undecanoic, and lauric acids together with several saturated and unsaturated alcohols. The use of (S)-(-)-2-methylbutanol afforded an optically pure ester. The heterogenized lipase is capable of inducing asymmetry during esterification with a prochiral alcohol. Both the entrapped lipase and the immobilized rhodium catalysts can be recovered simply by filtration and recycled in further runs without loss of catalytic activity.  相似文献   

6.
The reaction of N-nitro-O-(4-nitrophenyl)hydroxylamine (1) with conc. H2SO4 affords 4-nitropyrocatechol and that with conc. sulfonic acids (RSO3H where R = Me, CF3) affords 2-hydroxy-5-nitrophenyl-R-sulfonates in yields of 80?C85%. These reactions are assumed to proceed through an intermediate (phenoxy)oxodiazonium ion [NO2C6H4O-N=N=O]+, which eliminates the N2O molecule to form the aryloxenium ion [NO2C6H4O]+. The latter reacts with acid anions at the ortho-carbon atom of the phenyl ring. The thermodynamical parameters of the elementary reactions resulting in the formation of the (phenoxy)oxodiazonium ion [NO2C6H4O-N=N=O]+ and aryloxenium ion [NO2C6H4O]+ were calculated in the B3LYP/6?311+G(d) study of the combined molecular system (nitrohydroxylamine 1 + [H3SO4]+). The reaction of nitrohydroxylamine 1 with aqueous solutions of strong acids (??70% H2SO4, CF3SO3H) affords mainly 4-nitrophenol. It appears that the mechanism of this reaction does not involve the formation of the aryloxenium ion.  相似文献   

7.
Sulfonic acid chlorides react with hydroxylamine to form predominantly N-substituted products [1].Sulfonic acid fluorides react with hydroxylamine by formation of N-mesylhydroxylamine, hydrazinium(1+) and hydroxylammonium mesylate [2]. RfSO2F (Rf = C4F9, C8F17) reacting in the same way as CH3SO2F [3], the latter is considered to be a model compound for this kind of reactions.
The formation of the hydrazinium salt is explained by the occurrence of O-mesylhydroxylamine as intermediate. We conclude from the yields that sulfonic acid fluorides react with hydroxylamine to form predominantly O-substituted products.In order to prepare this O-derivative the reaction between O-(trimethylsilyl)hydroxylamine and mesylfluoride was carried out. Hydrazinium mesylate could be isolated also, which indicates the formation of the instable O-derivative.The reactions of methylhydrazines and silylated hydrazines with mesylfluoride lead to mesylated products as expected.  相似文献   

8.
Three new metal coordination complexes, namely, [Co ( DPE )(H2O)4]( DPE )( BS )2 ( 1 ), [Co ( DPE )2(H2O)4]( ABS )2 ( 2 ), [Co ( DPE )(H2O)4]( MBS )2(CH3OH)2 ( 3 ) [ DPE = (E)-1,2-di (pyridin-4-yl) ethene, BS = phenyl sulfonic acid, ABS = p-aminobenzene sulfonic acid, MBS = p-methylbenzene sulfonic acid] were obtained under hydrothermal conditions. Complexes 1 - 3 were structurally characterized by X-ray single-crystal diffraction, powder X-ray diffraction and IR. Complexes 1 and 3 exhibit a one-dimensional chain structure, and complex 2 does a zero-dimensional one. These three complexes further generate a three-dimensional supramolecular architecture via strong hydrogen bonding interactions and packing interactions. These three metal coordination complexes show high catalytic performance for green synthesis of a variety of 3,4-dihydropyrimidin-2(1H)-ones through the Biginelli reactions, which show several advantages such as excellent yields, short reaction times, eco-friendly synthesis conditions, and simple isolated workup procedure. Interestingly, the order of catalytic activities for these catalysts is the following: 3 > 1 > 2 , which can be ascribed to the acidities and hydrophobic interactions of phenyl sulfonate groups.  相似文献   

9.
The use of Montmorillonite K-10 as a convenient acid component in the Nicholas reaction is described. Its use permits functional selectivity, inter- and intramolecular reactions and convenience in the experimental conditions. A four-step one-pot [Co2(CO)6-acetylene complex formation, THP-removal, cyclization and complex cleavage] process permits the direct synthesis of 2-ethynyl-tetrahydrofuran from 6-(tetrahydro-2H-pyran-2-yloxy)hex-1-yn-3-ol.  相似文献   

10.
The reactions of -aminopropionic acid N"-acylhydrazides with aromatic and heterocyclic aldehydes and acetone afford compounds that exist in solutions predominantly as mixtures of 2-substituted 3-acylaminotetrahydropyrimidin-4-ones (ATHP) and tautomeric Schiff"s bases. These compounds in the crystalline state probably have structures of ATHP. The ratio of tautomers depends on the type of substituent in the aromatic ring and solvent. The reactions of 2-aryl-3-benzamidotetrahydropyrimidin-4-ones with carboxylic or sulfonic acid chlorides afford derivatives of 1-acyl- and 1-tosyl-3-benzamidotetrahydropyrimidin-4-ones.  相似文献   

11.
Attempts to prepare dipeptide compounds from organometallic N-substituted amino acids are reported. Condensation of pyridine-2-carboxaldehyde, α-amino acids (H-l-Ala-OH or H-l-Asp-OH) and W(CO)4(pip)2 leads to formation of W(CO)4(pyca-Et) (1) (pyca refers to the α-diimine fragment, C5H4NCHN) following decarboxylation of one or two equivalents of CO2. This decarboxylation does not occur for β-alanine or GABA (γ-aminobutyric acid). Coupling of [Hpip][W(CO)4(pyca-β-Ala-O)] (2) or [Hpip][W(CO)4(pyca-GABA-O)] (3) to amino acid esters, H-l-Ala-OEt, or H-l-Val-OMe, using the standard 1,3-dicyclohexylcarbodiimide (DCC), 1-hydroxybenzotriazole (HOBt) procedure produced four new dipeptide compounds, 4-7. The reactions proceed in good yield and compounds were characterized spectroscopically. The dipeptide complex, W(CO)4(pyca-Ala-Ala-OMe) (8), was prepared by reaction of W(CO)4(pip)2 with H-l-Ala-l-Ala-OMe and pyridine-2-carboxaldehyde. In addition the molecular structure of W(CO)4(pyca-β-Ala-Val-OMe) (5) is reported.  相似文献   

12.
Ca2+ complexation by both sulfate and selenate ligands was studied by CE. The species were observed to give a unique retention peak as a result of a fast equilibrium between the free ions and the complexes. The change in the corresponding retention time was interpreted with respect to the equilibrium constant of the complexation reaction. The results confirmed the formation of CaSO4(aq) and CaSeO4(aq) under our experimental conditions. The formation data were derived from the series of measurements carried out at about 15, 25, 35, 45 and 55°C in 0.1 mol/L NaNO3 ionic strength solutions, and in 0.5 and 1.0 mol/L NaNO3 ionic strength solutions at 25°C. Using a constant enthalpy of reaction enabled to fit all the experimental data in a 0.1 mol/L medium, leading to the thermodynamic parameters: ΔrG0.1M(25°C)=?(7.59±0.23) kJ/mol, ΔrH0.1 M=5.57±0.80 kJ/mol, and ΔrS0.1 M(25°C)=44.0±3.0 J mol?1 K?1 for CaSO4(aq) and ΔrG0.1 M(25°C)=?(6.66±0.23) kJ/mol, ΔrH0.1 M=6.45±0.73 kJ/mol, and ΔrS0.1 M(25°C)=44.0±3.0 J mol?1 K?1 for CaSeO4(aq). Both formation reactions were found to be endothermic and entropy driven. CaSO4(aq) appears to be more stable than CaSeO4(aq) by 0.93 kJ/mol under these experimental conditions, which correlates with the difference of acidity of the anions as expected for interactions between hard acids and hard bases according to the hard and soft acids and bases theory. The effect of the ionic medium on the formation constants was successfully treated using the Specific ion Interaction Theory, leading to significantly different binary coefficients mol/kg?1 and mol/kg?1  相似文献   

13.
The complexes of glycine, -alanine, and -alanine with (S)-[N-(N-benzylprolyl)amino] benzophenone formed by Ni(II) and Cu(II) ions and Schiff bases enter into different nucleophilic and electrophilic reactions with the formation of diastereoisomeric complexes which decompose into proteinogenic and nonproteinogenic L-amino acids with a high chemical yield and elevated optical purity (70–90%). Optically pure amino acids can be obtained from diastereoisomerically pure complexes after the complexes are separated by recrystallization of the mixture of diastereoisomeric complexes formed. A new type of interphase catalysts of C-alkylation of achiral Schiff bases was proposed. The catalysts are positively charged Ni(II) and Cu(II) complexes of Schiff bases of chiral diamines. In some cases, these complexes have a higher activity and capacity to execute asymmetric alkylation than traditional chiral interphase catalysts based on cinchonidine.Based on materials in the section report by Yu. N. Belokon' to the 7th European Symposium on Organic Chemistry, ESOC-7.A. N. Nesmeyanov Institute of Organoelemental Compounds, Russian Academy of Sciences, 117813 Moscow. Translated from Izvestiya Akademii Nauk, Seriya Khimicheskaya, No. 5, pp. 1106–1127, May, 1992.  相似文献   

14.
The reactions of the potentially tridentate Schiff bases 2-[(2-hydroxyphenyl)iminomethyl]phenol (H2ono) and 2-(2-aminobenzylideneimino)phenol (H3onn) with trans-[ReOBr3(PPh3)2] were studied, and the complexes [ReIIIBr(PPh3)2(ono)] (1) and [ReVBr(PPh3)2(onn)]Br (2) were isolated. In 1ono acts as a dianionic tridentate ligand, and in 2onn is coordinated as a tridentate trianionic imido-imino-phenolate. The complex [ReI(CO)3(ons)(Hno)] was isolated from the reaction of [Re(CO)5Br] with 2-[(2-methylthio)benzylideneimino]phenol (Hons; Hno = 2-aminophenol), with ons coordinated as a bidentate chelate with a free SCH3 group. These complexes were characterized by X-ray crystallography, NMR and IR spectroscopy.  相似文献   

15.
Abstract

The reaction of anthracene, 9-phenylanthracene and some 9-alkylanthracenes with dioxan-SO3 has been studied. Anthracene yields the 1-, 2- and 9-sulfonic acid in a ratio of 24 : 6 : 70. 9-Phenyl- and 9-neopentyl-anthracene both yield a mixture of the 4- and 10-sulfonic acids in a ratio of 33 : 67 and 15 : 85 respectively. Unexpectedly, 9-methylanthracene yields, in a more rapid reaction, exclusively 9-anthrylmethanesulfonic acid (1, R[dbnd]H) 9-Alkylanthracenes of which the alkyl group contains at least one α-H yield as main product (s) the α-sulfonic acids 1 and/or (depending on the further structure of the alkyl group) the sulfonic acids 24.  相似文献   

16.
17.
Abstract

Three pathways were observed in the reactions of Schiff bases of Thiohydrazides with P(NR2)3. (a) MeS-R2N exchange: MeS-C([dbnd]S)-NHN[dbnd]CHPh (1) reacted with P(NR2)3 led to new Schiff bases, R2N-C([dbnd]S)-NHN[dbnd]CH = Ph (2). (b) Cleavage of C[dbnd]S bond and the formation of P[dbnd]S bond: H2N-C([dbnd]S)-NHN[dbnd]CH = Ph (3) reacted with P(NR2)3 gave rise to the thiophosphoric amide. (Et2N)2P([dbnd]S)-NH-CH = N-N[dbnd]CH-Ph (4). (c) Formation of thiadiazole and triazole: Schiff bases 2a and H2N(MeS)C = N-N[dbnd]CH-Ph (6) reacted with P(NR2)3 respectively and produced 5-dimethylamino-2-phenyl-2,3-(2H)-1,3,4-thiadiazole (5) and 5-methylthio-2-phenyl-2,3-(2H)-1,3,4-triazole (7).  相似文献   

18.
Synthesis of a series of novel fused tetracyclic mono- and diazaindolo[2,1-c][1,4]benzoxazine heterocyclic compounds 3a-o has been achieved in a two-step one-pot reaction set up starting from commercially available or easily accessible inputs. For example, reaction of di-lithiated (N-Boc)-2-amino-3-methylpyridine Li2-2a with Weinreb amide of 2-(2,4-difluorophenoxy)-2,2-dimethylacetic acid 1a, followed by TFA treatment furnished the tetracyclic compound 3a, which is essentially a fusion of 7-azaindole and 3,4-dihydro[1,4]benzoxazine, in 70% isolated yield. A competitive elimination by-product 4a was also observed (24% isolated yield) in this case. Based on our initial results, a structural basis and molecular mechanism have been suggested to explain these two parallel reactions. Consequently, with appropriate structural tuning of 1, formation of the individual products can be controlled.  相似文献   

19.
The possibility of reactions between trans-[OsO2Cl2L2] (L = PPh3, AsPh3, SbPh3) osmium(VI) complexes and glacial acetic acid to give osmium(IV) compounds of general formula [Os2(-O)(-O2CCH3)2Cl4(L)2] was studied.  相似文献   

20.
The effect of Nb as a support modifier on the NiMo6/Al2O3–Nb2O5(x) (x?=?0, 1, 4, and 8?wt% Nb) catalysts was studied. The supports were prepared by one-pot coprecipitation from soluble precursors. The XRF analysis of the catalysts showed that the contents of Mo and Ni increased slightly with the presence of Nb. Micropore area and pore volume augmented importantly with Nb content, resulting in pore diameters between 5.3 and 9.3?nm. XPS analysis showed that the presence of Nb decreases the active metal–support interaction, improving the Mo and Ni sulfidation degree. The Raman spectra of sulfided catalysts suggested an increase in the number of layers of MoS2 in the presence of Nb. Generally, the thiophene HDS activity at normal pressure of sulfided NiMo6/Al2O3–Nb2O5(8) was greater than that of the sulfided catalysts with x?=?0, 1, and 4?wt% Nb, which can be attributed to the Nb promotion that would have an effect on the type of active site (Brønsted or Lewis acidic sites), since the number of sites by CO chemisorption for sulfided NiMo6/Al2O3–Nb2O5(x) did not show correlation with the catalytic activity. The high-pressure HDS activity of dibenzothiophene was also greater in the presence of Nb, and the hydrogenation route was preferred for the Nb-promoted solid, while the unpromoted one showed a larger yield of direct desulfurization 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