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 15 毫秒
1.
Zinc-mediated facile and efficient chemoselective S-alkylation of 5-aryl 1,3,4-oxadiazole-2-thiols in the presence of a catalytic amount of tetra butyl ammonium iodide was described. The reaction was performed under neutral conditions. The chemoselectivity of the alkylation was confirmed by NMR spectroscopy and x-ray crystallography.  相似文献   

2.
Reactions of poststerone with methyl iodide, allyl bromide, and propargyl bromide in lithium–ammonia solution resulted in its completely stereoselective 7α-alkylation accompanied by reduction of the 20-oxo group with formation of equimolar amounts of 20R- and 20S-hydroxy derivatives. The reaction of poststerone with excess allyl bromide afforded 7,7-bis-allyl 20R- and 20S-alcohols at a ratio of 3: 1. The reduction of the 20-oxo group in the alkylation of poststerone with excess propargyl bromide led to the formation of an equimolar mixture of Δ8(14)- and Δ8(9)-isomers with exclusive S configuration of C20.  相似文献   

3.
The selective reductive homo-coupling polymerization of aromatic diisocyanates via one-electron transfer promoted by samarium iodide in the presence of hexamethylphosphoramide (HMPA) produced the corresponding poly(oxamide)s ( 1 ) nearly quantitatively. The ob-tained polymers were insoluble in common organic solvents. The alkylation of 1 with methyl iodide or allyl bromide in the presence of potassium tert-butoxide provided the highly soluble alkylated polymer in good yields. In either case, the alkylation was almost complete, and both N-and O-alkylation proceeded. The ratio of N-and O-methylation was found to be 64 : 36 by 1H-NMR spectrum, and that of N- and O-allylation was 3 : 1 from 13C-NMR analysis. The present polymerization system could be applied to a variety of diisocyanates, including diphenylmethanediisocyanate (MDI), tolylene 2,6-diisocyanate (TDI), 2,6-naphthyl diisocyanate (NDI) and o-tolidine diisocyanate (TODI). The molecular weights of the polymers were estimated by GPC and found to be 2000–9000. The TGA measurement of the corresponding polymers showed Td10 at 248–320°C. © 1995 John Wiley & Sons, Inc.  相似文献   

4.
Trimethylsilyl and DBU nitronates derived from nitromethane and nitropropanes do not undergoC-alkylation by episulfonium (ESI) or thiophanium (TPI) cationic intermediates. The above-mentioned derivatives of 1-nitropropane react with ESI and TPI to give the corresponding products ofO-alkylation of 1-chloro-1-oxyiminopropane. The reactions of DBU nitronates derived from nitromethane and 2-nitropropane with ESI or TPI proceed asO-alkylation followed by standard fragmentation of the initially formed nitronate intermediates to give methyl (4-tolylthio)acetate or methyl 2,2-dimethyl-3-methyloxy-(4-tolylthio)butanoate respectively. Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 3, pp. 492–498, March, 1999.  相似文献   

5.
A synthesis of novel bis(aminopyrazoles) by the reaction of hydrazine hydrate with the appropriate bis(2‐cyanoketene‐S,N‐acetals) was reported. The latter compounds were prepared by treatment of bis(cyanoacetamides) with phenyl isothiocyanate in KOH/EtOH and subsequent alkylation with methyl iodide. The utility of bis(2‐cyanoketene‐S,S‐acetals) as building blocks for novel bis(aminopyrazoles) was also investigated.  相似文献   

6.
Novel ketene N,S‐acetals 3 were readily prepared by the reaction of cyanoacetamide or cyanothioacetamide with phenylisothiocyanate in the presence of potassium hydroxide, followed by alkylation of the produced salts with methyl iodide. The reaction of compounds 3 with hydrazines afforded different substituted pyrazoles 6.  相似文献   

7.
Alkylation of 4,5-disubstituted 4H-1,2,4-triazole-3-thiols with methyl chloroformate and ethyl chloroacetate chemoselectively afforded the corresponding S-alkyl derivatives, whereas the alkylation of 5-benzyl-4-phenyl-4H-1,2,4-triazole-3-thiol with methyl 3-bromopropanoate gave an inseparable mixture of S- and N-alkylation products. Hydrazinolysis of S-(5-benzyl-4-phenyl-4H-1,2,4-triazol-3-yl) methyl carbonothioate involved anomalous cleavage with formation of the initial 4,5-disubstituted 1,2,4-triazole and methyl hydrazinecarboxylate.  相似文献   

8.
The polarographic behavior of four S-methylthiophenium tetrafluoroborates, I-IV, has been investigated in Britton-Robinson aqueous buffers. The compounds under study were the S-methylbenzo[b]thiophenium (I), S-methylnaphtho[2,3-b]thiophenium (II), S-methyldibenzothiophenium (III), and S-methylnaphtho[2,1-b]thiophenium (IV) tetrafluoroborates which were obtained from the parent sulfur heterocycles by their reaction with methyl iodide and silver tetrafluoroborate in 1,2-dichloroethane, II and IV being new compounds. Compounds I-IV give two one-electron reduction waves and the more positive wave appears to be associated with an electron transfer from the electrode to the LUMO of the substrate as indicated by the experimental E1/2 vs calculated LUMO values plot (Figure 2). The rates of demethylation of the ions I-IV with the iodide ion acting as the nucleophile have been determined in the Britton-Robinson buffer (pH 6)-acetonitrile solution at 24°. The results of the polarographic study and of the kinetic studies with the iodide ion confirm the high reactivity of S-methylthiophenium ions which originates in the antiaromatic destabilization of these structures.  相似文献   

9.
《Tetrahedron: Asymmetry》2005,16(19):3160-3169
Representative (1S)- and (4S)-alkyl-1,4-dihydropyrazino[2,1-b]quinazoline-3,6-dione lactim ethers were regio- and diastereoselectively alkylated after metallation to give the corresponding 1,4-trans-isomers with retention of the stereocentres. The results were compared with the previously studied lactams. The (1R,4S)-dialkyl derivatives obtained by C(1)-alkyation with bifunctional reagents were lately cyclized to complex polycyclic compounds through a second N(2)-alkylation promoted by sodium iodide.  相似文献   

10.
《合成通讯》2013,43(14):2245-2253
ABSTRACT

The novel 2-thioxohydantoin ketene S,S-acetals 3a,b were prepared by the reaction of 2-thioxohydantoins 1a,b with carbon disulfide, followed by alkylation with methyl iodide of the produced sodium dithiolate salts 2a,b. The reaction of compounds 3a,b with nucleophiles afforded different fused methylsulfanylimidazoles.  相似文献   

11.
We have carried out a computational study of the reactive properties of chlorooxirane, the metabolically produced epoxide of vinyl chloride that is believed to be a direct-acting carcinogenic form of this molecule. An ab initio SCF-MO procedure (GAUSSIAN 70) was used to compute the energy requirements for stretching the C? Cl and both C? O bonds (SN 1 reactivity) and to determine the course of the epoxide's possible SN 2 reactions with ammonia, taken as a model for nucleophilic sites on DNA. The epoxide was assumed to be protonated; both the oxygen- and chloro-protonated forms were considered. At each step along the various reaction pathways, the structure of the system was reoptimized. For the oxygen-protonated epoxide, the C1? O bond has a significantly lower energy barrier to stretching than does the C2? O. (The carbon bearing the chlorine is designated C1.) However, both are very much higher than that of the C? Cl bond in the chloro-protonated form, confirming our earlier finding of the relative weakness of this bond. In the SN 2 processes involving ammonia, intermediate complexes are formed with both carbons of the oxygen-protonated epoxide, the C2-complex being the more stable. However, the most stable ammonia complex occurs at C1 of the chloro-protonated epoxide. Our calculated results, both the energies and also the geometry changes, allow us to propose two possible mechanisms for the formation of the 7-N-(2-oxoethyl) derivative of guanine that has been observed to be the major in vivo DNA alkylation product of vinyl chloride and has been suggested as possibly being responsible for its carcinogenicity. One of these mechanisms is SN 1 and starts with the chloro-protonated epoxide; the other is SN 2 and involves the oxygen-protonated form.  相似文献   

12.
The synthetic utility of 2-mercaptonicotinonitriles 3 and 4, as well as 2-mercapto-4-oxo-6-phenyl-1,4-dihydropyrimidine-5-carbonitrile 20 as building blocks for novel bis- and poly(pyridines), along with poly(pyrimidines) via alkylation with the corresponding bis- and poy(halo) compounds was investigated. Spectroscopic and theoretical studies confirmed the S-alkylation rather than the N-alkylation.  相似文献   

13.
A mild phase-transfer catalytic reaction has been conducted to O-alkylate the acidic functions of a lignite humic acid (HA), using tetrabutylammonium hydroxide as the phase-transfer catalyst. The HA acidic functional groups were made to react, in tetrahydrofuran, by nucleophilic substitution with several alkyl halides—methyl iodide, and ethyl, propyl, and butyl, and benzyl bromide. The occurrence of the O-alkylation reaction was assessed by elemental analysis and 1H NMR, CPMAS 13C NMR, and FTIR spectroscopy. Bonding of alkyl groups increased the carbon and hydrogen content and the H/C ratios of all the humic reaction products. Increased nitrogen in the reaction products suggested incomplete removal of the phase-transfer catalyst after purification of the alkylated HA. 1H NMR and CPMAS 13C NMR spectra of alkylated products provided evidence of the successful occurrence of the alkylation reactions. Infrared spectra confirmed the NMR results, revealing the characteristic absorption of newly formed alkyl and aryl ethers and esters in the alkylated products and C–H stretching in the aromatic ring of the benzylated derivative. These findings indicate that humic matter can be successfully alkylated with several different alkylating groups by catalysed phase-transfer reaction. This O-alkylation reaction has the advantage of being mild, versatile, and high-yielding compared with traditional methylation reactions applied to HA. The possibility of introducing different alkyl groups into the HA by a mild phase-transfer reaction may become useful by enabling improved fractionation of humic supramolecular associations and further understanding of the molecular nature of humic substances.  相似文献   

14.
Complementary Diastereoselective Cobalt Methylations of the Vitamin-B12 Derivative Cobester Treatment of heptamethyl cob(I)yrinate ( 2 ) in toluene/tetrahydrofurane (ca. 4:1) with methyl p-toluenesulfonate under exclusion of O2 and with protection from light leads to the selective formation of the heptamethyl Coβ-methylcob(III)yrinate (perchlorate 1b ) in 75% yield. In contrast, methylation of 2 with methyl iodide under the same conditons results in the isomeric heptamethyl Coα-methyulcob(III)yrinate (perchlorate 1a ) in 73% yield, besides 7% of 1b . This complementary diastereoface-selectivity of the methylation at the Co-center results from alkylation under kinetic control and apparently involves two mechanistically distinct alkylation processes. A radical mechanism is considered to account for the stereochemically unusual outcome of the reaction with methyl iodide.  相似文献   

15.
Sequence determination by partial methanolysis and fast atom bombardment (FAB) mass spectrometry of peptides containing cysteine and methionine was investigated. Cysteine-containing peptides require methylation of the sulphydryl group by methyl iodide to give a stable S-methylcysteinyl residue prior to partial methanolysis and mass spectrometry. Methionine-containing peptides undergo partially a methylation on sulphur during methanolysis, with formation of an S-methylsulphonium ion which under FAB conditions is extracted from the matrix and eliminates methyl sulphide in the gas phase. The presence of additional peaks due to chemical modifications or gas-phase fragmentations, however, does not interfere with the sequence information of the spectra.  相似文献   

16.
The Stereoselectivity of the α-Alkylation of (+)-(1R, 2S)-cis-Ethyl-2-hydroxy-cyclohexanecarboxylate In continuation of our work on the stereoselectivity of the α-alkylation of β-hydroxyesters [1] [2], we studied this reaction with the title compound (+)- 2 . The latter was prepared through reduction of 1 with baker's yeast. Alkylation of the dianion of (+)- 2 furnished (?)- 4 in 72% chemical yield (Scheme 1) and with a stereoselectivity of 95%. Analogously, (?)- 7 was prepared with similar yields. Oxidation of (?)- 4 and (?)- 7 respectively furnished the ketones (?)- 6 (Scheme 3) and (?)- 8 (Scheme 4) respectively, each with about 76% enantiomeric excess (NMR.). It is noteworthy that yeast reduction of rac- 6 (Scheme 3) is completely enantioselective with respect to substrate and product and gives optically pure (?)- 4 in 10% yield, which was converted into optically pure (?)- 6 (Scheme 3). The alkylation of the dianionic intermediate shows a higher stereoselectivity (95%) from the pseudoequatorial side than that of 1-acetyl- or 1-cyano-4-t-butyl-cyclohexane (71% and 85%) [9] or that of ethyl 2-methyl-cyclohexanecarboxylate (82%). The stereochemical outcome of the above alkylation is comparable with that found in open chain examples [1] [2]. Finally (+)-(1R, 2S)- 2 was also alkylated with Wichterle's reagent to give (?)-(1S, 2S)- 9 in 64% yield. The latter was transformed into (?)-(S)- 10 and further into (?)-(S)- 11 (Scheme 5). (?)-(S)- 10 and (?)-(S)- 11 showed an e.e. of 76–78% (see also [11]). Comparison of these results with those in [11] confirmed our former stereochemical assignment concerning the alkylation step.  相似文献   

17.
The main products of the reaction of terephthalonitrile dianion disodium salt with ω-X-alkyl bromides (2-(2-bromoethyl)-1,3-dioxane, 5-bromovaleronitrile, ethyl 6-bromohexanoate) in liquid ammonia are the corresponding 4-(ω-X-alkyl)benzonitriles. Similar reactions of benzonitrile radical anion sodium salt lead to ω-X-alkylbenzenes. In both cases the formation of products is due to selective ipso-alkylation of anionic forms that indicates the nucleophilic activity of terephthalonitrile dianion and benzonitrile radical anion in these reactions and the realization of alkylation via S N 2 mechanism.  相似文献   

18.
《Tetrahedron: Asymmetry》2001,12(7):983-986
Polymer-supported cinchona alkaloid salts with different spacers were used as phase-transfer catalysts in the asymmetric C-alkylation of N-diphenyl methylene glycine tert-butyl ester. Various catalysts and alkylation conditions were studied, the best result being 81% e.e. with cinchoninium iodide bound to polystyrene with a four-carbon spacer.  相似文献   

19.
A method for the synthesis of carboxylic acid derivatives containing one or two —CH2CH n (Me)CH n+1CH2— fragments (n = 0, 1) was developed. The method is based on the alkylation of (di)alkyl malonates, cyanoacetates, and acetoacetates with acyclic prenyl halides in ionic liquids, 1-butyl-3-methylimidazolium hexafluorophosphate and tetrafluoroborate. For the ambident ethyl acetoacetate anion, the reactions with prenyl halides devoid of a double bond in the allylic position relative to the halogen atom carried out in the ionic liquids give mixtures of C- and O-alkylation products, while in the case of allylic prenyl halides, only C-alkylation products are formed. The reactions of ethyl 2-geranylmalonate and 2-geranylacetoacetate with bromocyclohexane and 1-chloro-3-dimethylaminopropane in ionic liquids provided derivatives of pharmacologically active geranylacetic acids. The product yields are higher than those in molecular organic solvents. The ionic liquids were recovered and reused in the alkylation.  相似文献   

20.
 5-Isopropyl-6-naphthyl uracil and 5-isopropyl-6-naphthyl-2-thiouracil were alkylated to give N-1-(ethoxymethyl and methylthiomethyl) uracil and S2-cyclohexyl-thiouracil, respectively. 5-Ethyl-6-naphthyl uracil and 5-ethyl-6-naphthyl-2-thiouracil afforded N-1-(ethoxymethyl, methoxy-methyl, methylthiomethyl, acetoxyethoxy methyl and hydroxyethoxy methyl) uracil and S2-((2,2- diethoxyethyl), methoxycarbonylmethyl, ethoxycarbonylpropyl, methylthiomethyl, ethoxymethyl, methyl and cyclohexyl)-thiouracil upon alkylation.  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