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.
From results of oxidation, protonation, and alkylation of the products arising in one- or two- electron reduction of m-tolunitrile with sodium in liquid ammonia followed a conclusion that these products are respectively anion-radical of the compound and 3-methyl-1-cyano-2,5-cyclohexadienyl anion. The reaction of both reduction products with alkyl halides gives rise to compounds of ipso-alkylation with respect to cyano group: the corresponding alkyltoluenes and 1-alkyl-3-methyl-1-cyclohexadienes. The ratio of these products depends on the structure of alkyl halide. The possibility to prepare selectively m-alkyltoluenes by reaction of the product of two-electron reduction of m-tolunitrile with alkyl halides was demonstrated.  相似文献   

2.
Birch's reductive alkylation of biphenyl-4-carbonitrile (1) provides alkylated 1,4-dihydroderivatives of various structural types: 4-alkyl-4-phenylcyclohexa-2,5-dienone, 1,4-dialkyl-4-phenylcyclohexa-2,5-dienecarbonitrile (with the same or different alkyl fragments), and 4-(1-alkylcyclohexa-2,5-dienyl)benzonitrile. Each of these products become dominant depending on the nature of long-living anionic form generated from 1, namely, the stable product of two-electrons reduction – dianion (12?); 1-alkyl-4-cyano-1-phenylcyclohexa-2,5-dien-4-yl anion (1-Alk1), originated due to the alkylation of dianion 12? at the position 1 of biphenyl moiety; or 1-(4-cyanophenyl)cyclohexa-2,5-dien-1-yl anion (1-H4’), being the product of dianion 12? protonation at position 4′ by protonating reagent (MeOH or NH4Cl). The orientation of alkyl fragment incorporation into biphenyl-4-carbonitrile scaffold is in agreement with calculated electronic structure of the anionic species under investigation. The dominating type of their reactivity towards alkyl halides proved to be nucleophilic (SN2 mechanism).  相似文献   

3.
Homarine (1-methyl-2-carboxypyridinium ion) and trigonelline (1-methyl-3-carboxy-pyridinium ion) are reduced at mercury electrodes in alkaline solutions by a one-electron transfer to the pyridinium ring. Analysis of polarographic wave shapes using deviation-pattern recognition, peak widths of cyclic voltammograms, analysis of the products of bulk electrolyses and dependences of E1/2 and Ep on scan rate, pH and concentration indicated that these reductions proceed by an EC2 mechanism, where the second-order chemical reaction is a homogeneous dimerization. The zwitterionic form of the pyridiniumcarboxylic acid, which predominates in alkaline solutions, undergoes reduction to form a pyridinyl anion radical. Two radicals then couple to yield a dimer dianion which is subject to protonation. The dimer can be oxidized at mercury electrodes.  相似文献   

4.
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   

5.
The electrochemical reduction of the cyclohex-2-enones 1a–1e (mercury cathode, CH3CN, Bu4NBF4) was studied by means of cyclic voltammetry, d.c. polarography, coulometry and chemical product analysis. Compounds 1a–1c give a mixture of the hydrodimers 4 and 5 via formation of the radical anion 2 by an irreversible one electron transfer, followed by protonation and dimerization of the allylic radical 3 . The 6-halocyclohex-2-enones 1d and 1e exhibit two distinct reduction waves. The first corresponds to an irreversible two electron transfer with formation of the halide anion and the enolate anion 6 which gives 1b by protonation. The second wave corresponds to a quasi-reversible one electron transfer to 6 to afford the radical dianion 7 (Scheme 2).  相似文献   

6.
C. Schnieders  K. Müllen  W. Huber 《Tetrahedron》1984,40(10):1701-1711
Reduction of pyrene with alkali metals yields the corresponding dianion salts. The solvent, counterion and temperature must be carefully selected since side reactions such as protonation (e.g. in liquid ammonia) or cleavage of the etheral solvent occur readily. Moreover, the spectroscopic characterization of the dianion is complicated by rapid electron transfer processes. There is no experimental evidence for distorted dianion structures or for further reduction of pyrene toward a tetraanion. Knowledge of the ionic π-structures is essential for an understanding of reductive alkylation processes.  相似文献   

7.
The synthesis of 5,10-dideazaaminopterin by two independent routes is described. Condensation of the piperidine enamine of 4-p-carbomethoxyphenylbutyraldehyde ( 4 ) with ethoxymethylenemalononitrile followed by treatment of the resultant arylethylenaminomalononitrile ( 5 ) with methanolic ammonia produced 2-amino-3-cyano-5-p-carbomethoxyphenethylpyridine ( 6 ). Cyclization of the aminocyanopyridine with guanidine afforded 4-amino-4-deoxy-5,10-dideazapteroic acid ( 8 ). Coupling of the pteroate intermediate with glutamate yielded the target 5,10-dideazaaminopterin ( 10 ). Alternatively, reduction of 2,4-diamino-6-formyl-5-deazapteridine ( 11 ) with sodium borohydride gave the 6-hydroxymethyl compound 12 . Conversion to the bromide was followed by alkylation of dimethyl homoterephthalate to afford methyl 4-amino-4-deoxy-10-carbomethoxy-5,10-dideazapteroate ( 14 ). Decarboxylation with ester cleavage (sodium cyanide in dimethyl sulfoxide at 180°) also gave the diaminopteroic acid ( 8 ). 5,10-dideazaaminopterin ( 10 ) was an effective growth inhibitorof folate dependent bacteria, S. faecium and L. casei.  相似文献   

8.
Irradiation of 3,5-diphenyl- or 3-(p-tolyl)-5-phenyl-2-isoxazoline ( 12 and 13 , respectively) in benzene with a high-pressure mercury lamp yields 4,5-diphenyl- or 4-(p-tolyl)-5-phenyl-3-oxazoline ( 17 and 19 , respectively) and the β-amino-chalcones 18 or 20 in addition to benzaldehyde, benzonitrile and p-tolunitrile, respectively (scheme 6 and ‘Anmerkg.’ p. 2600). The 3-oxazolines 17 and 19 are formed by route a (scheme 8) via 3-phenyl- or 3-(p-tolyl)-2H-azirine ( 23 , R = H and CH3, respectively) and their photochemically rearranged successors, the nitrile methylides 24 , as intermediates. The discovery of this reaction has served as a basis for the quickly developing photochemistry of 3-aryl-2H-azirines [2] [24]. Photolysis of the 2-isoxazoline 13 in methanol leads to the formation of a mixture of syn/anti-p-tolyl trans-styryl ketoximes (syn/anti, trans- 30 ) and anti, cis- 30 , 2-(p-tolyl)-quinoline ( 29 ), the 4-hydroxymethylated derivative 32 of the latter (in small amounts), besides the β-aminochalcone 20 , benzaldehyde, p-tolualdehyde and p-tolunitrile (scheme 9). It could be shown that the stereoisomeric ketoximes 30 are photochemically interconvertible (scheme 12) and that at least one mechanism of formation of 2-(p-tolyl)-quinoline ( 29 ) is the photo-induced cyclisation of p-tolyl-cis-styryl ketoximes (cis- 30 ) (scheme 13). A tentative mechanism for the formation of p-tolual-dehyde is given in scheme 10; the crucial step is the protonation of p-tolunitrile methylide ( 24 , R = CH3) by methanol at the nitrile carbon atom, after which hydrolysis yields the aldehyde.  相似文献   

9.
A reaction of methyllithium with 3-cyano-6-methylpyridine-2(1 H)-thione followed by alkylation of the resulting 3-acetylpyridinethione, or a direct reaction of methyllithium with 3-cyano-6-methyl-2-(methylthio)pyridine, afforded 3-acetyl-6-methyl-2-(methylthio)pyridine. The ketone obtained was examined in bromination reactions under various conditions. Bromi-nation in methanol or chloroform, proceeding through the formation of sulfonium bromides, gave substituted 3-(bromoacetyl)pyridine. A reaction of 3-acetyl-6-methyl-2-(methyl-thio)pyridine with N-bromosuccinimide in CCl4 afforded N-(pyridinesulfenyl)succinimide. The bromo ketone was used for the synthesis of various heterocyclic compounds.  相似文献   

10.
The two-electron, two-proton reduction of alkynes to trans-alkenes has been studied computationally using the polarizable continuum regime to model liquid ammonia, the solvent in which such reductions are generally carried out. Two computational approaches have been used. In one, the energies of species (alkyne, radical anion, vinyl radical, vinyl anion, dianion, and alkene) that are implicated as possible participants in the reduction are obtained using high level ab initio single-point computations under the polarizable continuum model (PCM) conditions. In the other approach, the same species are surrounded by ten explicit ammonia molecules before undergoing the same single-point PCM analysis. It has been shown that the two methods provide nearly identical results in terms of relative energies. Other findings include the probable bent nature of the radical anion species in ammonia, the likelihood that the trans stereochemistry of the reduction is determined at the vinyl anion stage, and the elimination of a dianion as a possible species that determines the stereochemical result. Various observations relating the solvent effects of ammonia are made relative to known gas-phase properties of the species studied.  相似文献   

11.
The photoinitiated reactions of 2,4-pyridinedicarbonitrile ( 1 ) and benzophenone in neutral, acidic, and basic 3:1 2-propanol-water and the kinetics of disappearance of 1 , have been studied. Pyridinyl radical anion forms as an intermediate by an electron transfer. In acidic solution substitution of the cyano group in the 2 position and in the 2 and 4 positions with diphenylmethanol occurs. In neutral medium both substitution at the 2-position and reduction, in which the cyano group at the 4 position is replaced by hydrogen, are observed. In basic solution in which protonation of the radical anion is not likely, only reduction occurs. The rates of formation and relative yields of these products depends on the pH of the solution. A mechanism is discussed.  相似文献   

12.
Malononitrile reacted with the title compound to give 6-amino-5-cyano-2-(3,3-dicyano-2-methylallylidene-4-methyl-2H-pyran (3). Treatment of 3 with hot 80% sulfuric acid yielded 4,7-dimethyl-56-hydroxy-2(1H)quinolone. With concentrated aqueous sodium hydroxide, 3 gave 5-amino-3,6-dicyano-4,7-dimethyl-2(1H)quinolone and 5-amino-6-carbamoyl-3-cyano-4,7-dimethyl-2(1H)quinolone. The reaction of 3 with hydrochloric in acetic acid gave a mixture of 6-amino-3,7-dicyano-2,8-dimethyl-4-quinolizone and 3-cyano-4-methyl-6-(3,3-dicyano-2-methylallyl)-2-pyrone. Compound 3 also reacted with methylamine, butylamine and piperidine to give 8-amino-5-cyano-4-methyl-2-pyridone, 6-bulylamino-5-cyano-4-methyl-2-pyridone and 5-eyano-4-methyl-6-piperidino-2-pyridone respectively.  相似文献   

13.
Reactions of methyl 2-cyano-3-methyl-3-(p-substituted-phenyl)acrylates ( 1 ) with 1-phenyldiazoethane ( 2 ) produced the stable abnormal-addition 1-pyrazolines, AP, 4 [1]. The structure of cis-syn- 4a (X = NO2) was determined by X-ray crystallography. Thermal decomposition of 4 results in “true cycloreversion” to the starting materials. Stereochemistry and decomposition of the normal-addition 1-pyrazolines, NP , produced in situ from 2-cyano-3-(p-substituted-phenyl)acrylates and diazoalkanes, are also discussed.  相似文献   

14.
The kinetics of 4-nitrochlorobenzene reactions with substituted phenolates in the medium of N,N-dimethylacetamide was studied. The BrØnsted relation is fulfilled by substituted potassium phenolates: the nucleophilicity of phenolates increases with an increase in their basicity. The rate-limiting step in the reactions of 4-nitrochlorobenzene with substituted phenolates and potassium resorcinate is changed from the phenoxide anion to the phenoxide dianion. In the latter case, electron transfer from the resorcinate dianion with the generation of radical species can be responsible for the reaction rate.  相似文献   

15.
The reductive electrochemistry of substituted benzophenones in the aprotic room-temperature ionic liquid (RTIL) 1-butyl-1-methylpyrrolidinium bistriflimide occurs via two consecutive one-electron processes leading to the radical anion and dianion, respectively. The radical anion exhibited electrochemical reversibility at all time-scales whereas the dianion exhibited reversibility at potential sweep rates of >or=10 V s(-1), collectively indicating the absence of strong ion-paring with the RTIL cation. In contrast, reduction in 1-butyl-3-methylimidazolium bistriflimide is complicated by proton-transfer from the [Bmim] cation. At low potential sweep rates, reduction involves a single two-electron process characteristic of either an electrochemical, chemical, electrochemical (ECE) or disproportion-type (DISP1) mechanism. The rate of radical anion protonation in [Bmim] is governed by basicity and conforms to the Hammett free-energy relation. At higher potential sweep rates in [Bmim][NTf2], reduction occurs via two consecutive one-electron processes, giving rise to the partially reversible generation of the radical anion and the irreversible generation of the dianion, respectively. Also, the redox potentials for the reversible parent/radical anion couples were found to be a linear function of Hammett substituent constants in both RTIL media and exhibited effectively equivalent solvent-dependent reaction constants, which are similar to those for reduction in polar molecular solvents such as acetonitrile or alcohols.  相似文献   

16.
The major products of reactions of the terephthalonitrile radical anion with α,ω-dibromoalkanes Br(CH2)nBr (n = 3–5) were 4-(ω-bromoalkyl)benzonitriles. Analogous reactions of the terephthalonitrile dianion mainly yielded α,ω-bis(4-cyanophenyl)alkanes. Both transformations are convenient one-step routes to otherwise not easily accessible compounds that are valuable as versatile building blocks. The results of alkylation allow one to suggest that reactions of the dianion with intermediate 4-(ω-bromoalkyl)benzonitriles proceed more rapidly than those with the starting α,ω-dibromoalkanes. This was confirmed by competitive reactions of the dianion with 4-(ω-bromoalkyl)benzonitriles and the corresponding alkyl bromides. To explain such a ratio of the reaction rates, a mechanism was proposed for the reaction of the dianion with 4-(ω-bromoalkyl)benzonitriles. According to this mechanism, a charge transfer complex is a key reaction intermediate. Dedicated to the memory of Academician N. N. Vorozhtsov on the 100th anniversary of his birth. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 6, pp. 1069–1077, June, 2007.  相似文献   

17.
The rearrangement of 2-amino-3-carbethoxy-4-ethylfuro[3,2-b]pyridinium iodide in basic solution was studied. The reaction product is 2-oxo-3-cyano-4-ethyl-4H-furo[3,2-b]pyridine which was obtained also by alkylation with ethyl iodide and sodium hydride in dimethylformamide of 2-oxo-3-cyano-3H-furo[3,2-b]-pyridine or of p-nitrophenyl-3-acetoxypyridine-2-cyanacetate.  相似文献   

18.
The reaction of nucleophilic and non-nucleophilic bases wtih 2-carbamoyl-3-(γ-chloropropyl)-1-indenone ( 5 ) have been investigated. Condensation of γ-chlorobutyrophenone with malono-nitrile afforded α-cyano-β-(3-ehloropropyl)cinnamonitrile which was cyclized in concentrated sulfurie acid to produce 5 . Two other products obtained from the cyclization reaction were 2-carbamoyl-3-(γ-ehloropropylidene)-1-indanone ( 4 ) and α-carbamoyl-β-(3-chloropropyl)cinnam-amide. Treatment of a solution of 4 in ethyl acetate with piperidine resulted in cyclization of the γ-chloropropyl side chain to give 2-carbamoyl-3-cycIopropyl-1-indanone. The same compound was obtained in improved yield by the treatment of 4 or 5 with sodium hydroxide solution. The reaction of dirnethylamine with 5 in benzene gave initial Michael addition of the amine followed by internal alkylation of the carbanion so formed to yield 3a-dimethylamino-2,3,3a,8-tetrahydro-8-oxoeyclopent[a]indene-8a(lH)earboxamide ( 7a ). Similarly addition of ammonia, pyrrolidine, piperidine, benzenethiol, p-toluenethiol, 2-naphthalenethiol and nitromethane to the indenone I gave respective analogs of type 7 . Treatment of 5 with sodium cyanide in aqueous t-butyl alcohol resulted in a similar Michael addition followed by internal alkylation. In addition, cyclization between the nitrile and the carbamoyl functions occurred in the same step to give 2-oxo-4-imino-7,8-benzo-3-aza[3.3.3]-propellan-6-one ( 13a ). Hydrolysis of the iminopyrrolido ring in 13a to the corresponding suecin-irnide gave 2,4-dioxo-7,8-benzo-3-aza[3.3.3]propellan-6-one ( 13b ). Reactión of 13b with methyl iodide, allyl bromide, benzyl bromide, and diethyluminoethyl chloride afforded the corresponding N-alkylated products. A similar sequence starling with δ-ehlorovalerophenone led to 5,6-fused ring systems, including a [4.3.3]propellane. 2,9-Dioxo-4-methyl-7,8-benzo-3-aza[4.3.3]propell-4-ene was obtained by the reaction of 5 with acetone in dilute alkali.  相似文献   

19.
The alkylation and sulfonylation of 3-aryl-1-methyl-1,2,4-triazolin-5-ones (1) were studied with various alkyl halides and sulfonyl chlorides. The alkylation of 1 with methyl iodide and ethyl bromide afforded N-alkylated products, however with methyl 2-bromopropionate afforded O-alkylated products predominantly. The sulfonylation by methanesulfonyl chloride afforded a mixture of N-sulfonylated and O-sulfonylated products, while the sulfonylation by p-toluenesulfonyl chloride afforded mainly O-sulfonylated products.  相似文献   

20.
Three derivatives of 2-amino-6-methyl-5-(p-tolylsulfonamidopropyl)-4-pyrimidinol (I) with N - substituents on the sulfonamide group, namely bromoacetamidopropyl (XVI), m-bromoacetamidobenzyl (XXIVa), and p-bromoacetamidobenzyl (XXIVb), were synthesized as candidate active-site-directed irreversible inhibitors of thymidylate synthetase. The bromoacetamidopropyl derivative, (XVI), the p-bromoacetamidobenzyl derivative (XXIVb), and iodoacetamide showed irreversible inhibition of thymidylate synthetase, but XXIVa did not. Since iodoacetamide did inactivate the enzyme, but XXIVa did not, it cannot be ascertained whether XXIVb and XVI inactivate the enzyme by a random bimolecular mechanism or by the active-site-directed mechanism without evaluation of additional candidate inhibitors. Two synthetic routes were employed. The key intermediates for the bromoacetamidobenzyl sulfonamides (XXIV) were the corresponding nitrobenzyl sulfonamides (XXI); the latter were best prepared by reductive alkylation of 2-amino-5-aminopropyl-6-methyl-4-pyrimidinol (XXV) with a nitrobenzaldehyde followed by tosylation. The key intermediate for XVI was a toluenesulfonamide with a carbobenzoxyaminopropyl substituent on the nitrogen (XIV); the latter was synthesized via N-carbobenzoxy-N'-tosyl-1,3-diaminopropane (XI).  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