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 718 毫秒
1.
Abstract

Sequential tritylation, benzoylation and detritylation of D-glucose, followed by resolution of the crude product by chromatograpEy gave crystalline 1,2,3,4-tetra-O-benzoyl-α- (1) and β-D-glucopyranose (2). Compound 1, 2, and the corresponding methyl α-glycoside 5 were treated with dimethylaminosulfur trifluoride (methyl DAST) to give, respectively, the 6-deoxy-6-fluoro derivatives 3, 4, and 6. Crystalline 2,3,4-tri-O-benzoyl-6-deoxy-6-fluoro-α-D-glucopyranosyl chloride (10) could be obtained from either 3, 4, or 5 by reaction with dichloromethyl methyl ether in the presence of anhydrous zinc chloride. Silver trifluoromethanesulfonate-promoted reaction of 10 with methyl 2-O-(9) and 3-O-benzyl-4,6-O-benzylidene-α-D-glucopyranoside (8) gave the corresponding, (β-linked disaccharidës in high yield. Subsequent deprotection afforded the 6′-deoxy-6′-fluoro derivatives of methyl α-sophoroside (13) and methyl 6′ -deoxy-o′-fluoro-α-laminaribioside (16). Condensation of 8 and 9 with 6-O-acetyl-2,3,4-tri-O-benzyl-α-D-glucopyranosyl chloride in the presence of silver perchlorate was highly stereoselective and produced the α-linked disaccharidës 17 and 21, respectively, in excellent yield. Deacetylation of 17 and 21, followed by fluorination of the resulting alcohols 18 and 22 with methyl DAST and subsequent hydrogenolysis, gave 6′-deoxy-6′-fluoro derivatives of methyl α-kojibioside and methyl α-nigeroside 20 and 24, respectively.  相似文献   

2.
Diphenylmethane and fluorene were used as target molecules in an investigation of the effect of the geometry of aromatic molecules on the regioselectivity and rate of fluorination with 1-chloromethyl-4-fluoro-1,4-diazoniabicyclo[2.2.2]octane bis(tetrafluoroborate) (Selectfluor F-TEDA-BF4). In acetonitrile at 80 degrees C ring fluorination of diphenylmethane was accompanied by oxidation of the saturated carbon atom, while in trifluoroacetic acid only ring fluorination with an ortho-para regioselectivity of 1.8:1 was observed. Fluorene was converted in acetonitrile as well as in trifluoroacetic acid into 2- and 4-fluoro substituted products in the relative ratio of 2:1 and 1.2:1, respectively. The reactions in acetonitrile obey a simple rate equation: v = d[F-TEDA]/dt = k2 x [F-TEDA] x [Substrate] and the second order rate constants for the reactions in acetonitrile at 65 degrees C were determined; values of 0.6 x 10(-4) M-1 s-1 for diphenylmethane and 35.5 x 10(-4) M-1 s-1 for fluorene were obtained. The reaction rates for the various functionalisations of fluorene relative to those for diphenylmethane were found to be considerably influenced by the type of functionalisation. Relative rate factors (k(rel) = k2(fluorene)/k2(diphenylmethane)) with values between 59 for fluorination and 712 for chlorination were determined, while the corresponding data for the biphenyl/diphenylmethane pair were only slightly dependent on the type of functionalisation. A reaction pathway involving electron transfer, thus forming cation radical intermediates, was proposed as the main process in the case of fluorination of fluorene with F-TEDA-BF4.  相似文献   

3.
Expanding the repertoire of controlled radical fluorination techniques, we present a photosensitized unstrained C–C bond activation/directed monofluorination method using Selectfluor and 9-fluorenone. The reaction is amenable to the opening of multiple 1-acetal-2-aryl substituted rings to yield ω-fluoro carboxylic acids, esters, alcohols, and ketones with relative ease. Initial mechanistic insight suggests radical ion intermediates.  相似文献   

4.
A highly efficient one-step photocyanation reaction of pyrene was shown to proceed in oil-in-water emulsion systems. As a typical example, photoirradiation of pyrene in the presence of 1,4-dicyanobenzene and NaCN in a benzonitrile/water mixture (1/100, vol %) under vigorous stirring gave 1-cyanopyrene in a conversion yield of 83%, while an analogous reaction in an acetonitrile/water mixture (9/1, vol %) yielded this product in 61% yield. We evaluated the quantum yield of photocyanation in the oil-in-water emulsion system to be 0.17. Under optimum conditions, the quantum yield was improved to 0.68. Characteristics and possible mechanisms of the photocyanation reaction are discussed in detail.  相似文献   

5.
Electrochemical fluorination of anisole furnished 2-and 4-fluoroanisoles in a 3:1 ratio, guaiacol, and 4,4′-dimethoxydiphenyl ether. Phenylacetonitrile alongside the fluorination in the ring suffered the transformation of the cyano group into a trifluoromethyl. 4-Bromobenzamide was fluorinated to a high conversion mostly in the ring to afford predominantly 4-bromo-3,3,6,6-tetrafluoro-1,4-cyclohexadienecarboxamide. 4-Bromonitrobenzene in a low yield gave 4-bromofluoronitrobenzene and 3,4-dibromofluoronitrobenzene. 3-Bromo-nitrobenzene and 1,4-dichlorobenzene lid not undergo fluorination. In the course of the electrolysis of the 4-bromobenzamide and 4-bromonitrobenzene in anhydrous HF apart the fluorination occurred also the bromination of the substrates.  相似文献   

6.
Electrochemical fluorination of 2-methoxy-1,1,1-trifluoro-2-(F-methyl) octane gave the corresponding perfluorinated ether in 27% yield, along with cyclic by-products in 9%. A mixture of partly fluorinated tertiary amines, consisting of 1-dipropylamino-F-1-propene and 1-dipropylamino-2-hydryl-F-propane, did not afford a superior yield of tri(F-propyl)amine compared to the unfluorinated tripropylamine. 1-Diethylamino-2-(F-methyl)-F-1-pentene was fluorinated to give the corresponding F-tertiary amine in fairly good yields, together with 1-di(F-ethyl)amino-2-hydryl-2-(F-methyl)-F-pentane and their fragmented products. The study indicates that blocking of the α-carbon atom of an ether with F-methyl groups seems to reduce fragmentation, resulting in good yields of an unrearranged product. However, partial fluorination of a tertiary amine prior to electrochemical fluorination rather allows high yields of undesired by-products, as far as our experiments were concerned.  相似文献   

7.
We have found the new nucleophilic fluorination reaction of some halo- and mesylalkanes to the corresponding fluoroalkanes with KF in the presence of [bmim][BF4] under various reaction conditions. 2-(3-Methanesulfonyloxypropoxy)naphthalene (1) was used as a model compound to optimize this fluorination reaction. Whereas the fluorination of the mesylate 1 with KF in an organic solvent such as CH3CN at 100 degrees C occurred hardly even after 24 h, the same reaction in ionic liquids, [bmim][BF4], as a reaction solvent was completed within 1.5 h, affording the wanted product 2-(3-fluoropropoxy)naphthalene 2a (85%) together with the alkene byproduct 2c (10%). Very interestingly, however, the addition of water (5 equiv) completely eliminated the formation of the undesired alkene 2c and thus gave higher yield of 2a (92%, entry 2). The use of acetonitrile as a cosolvent did not affect the reactivity of the fluorination. The presence of a proper amount of cosolvent was rather desirable (94% yield of 2a). We performed fluorination reactions with other ionic liquids ([bmim][PF6], [bmim][SbF6], [bmim][OTf], and [bmim][N(Tf)2], and two other cosolvents, to find the optimal ionic liquid and cosolvent. Nine different compounds were examined, including the 10 g-synthesis of 2-(fluoromethyl)naphthalene in 93% of isolated yield.  相似文献   

8.
(+/-)-7beta,8alpha-Dihydroxy-9beta,10beta-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (DE-1) undergoes reaction with anhydrous HCl in dioxane to yield predominantly ( approximately 94%) a single chlorohydrin. This chlorohydrin was assigned structure 9, in which the chloro goup at C-10 is located cis to the C-9 hydroxyl group, on the basis of its (1)H NMR spectrum. This result is in contrast to the reaction of a diastereomeric benzo[a]pyrene 7,8-diol 9,10-epoxide (DE-2) with HCl, which yields only trans-chlorohydrin 8. The hydrolysis of cis-chlorohydrin 9 in 10:90 dioxane-water solutions yields the same ratio of tetrols ( approximately 89% cis/11% trans) as that formed by acid-catalyzed hydrolysis of DE-1. This result again contrasts with the hydrolysis of trans-chlorohydrin 8, which undergoes hydrolysis to give tetrols in a ratio different from that from acid-catalyzed hydrolysis of DE-2. A marked common ion rate depression in the hydrolysis of cis-chlorohydrin 9 is observed, which shows that hydrolysis proceeds via an intermediate carbocation that has a sufficient lifetime to be trapped by external chloride ion. The observation that DE-1 reacts with HCl to give mainly the cis-chlorohydrin is rationalized by quantum chemical calculations that suggest that the cis-chlorohydrin is more stable than the epimeric trans-chlorohydrin.  相似文献   

9.
A series of novel coumarin glycoside esters(1―9) was synthesized through the acylation reaction of 4-methylcoumarin-7-O-β-D-glucoside(11) with different long chain fatty acids catalyzed by lipase in organic medium. The acylation occurred regioselectively at the 6'-OH of glycosyl moiety. The enzymatic synthesis was optimized to achieve 54%―70% yield using immobilized lipase(Novozym 435, 10 mg/mL) as catalyst and acetone and pyridine(9:1, volume ratio, water content<1%) as solvent with an acyl donor/coumarin glycoside molar ratio of 10:1 at a temperature of 40―50 ℃ for 96 h. All the synthesized compounds were confirmed.  相似文献   

10.
Upon reaction of the 3',4'-unsaturated adenosine derivative 2 with N-iodosuccinimide (NIS) and thiophenol, an unexpected electrophilic hydrophenylsulfanylation proceeded to provide 4'-phenylsulfanylcordycepin 7 in 79% yield with the ratio 7a/7b = 6.6/1. A study of the reaction mechanism revealed that hydrogen iodide (HI) generated from NIS and PhSH acted as an active species. On the basis of a deuterium experiment using PhSD, initial protonation occurred at the β face of the double bond to furnish the β-π complex III, which underwent anti addition of PhSH as a major pathway. Nucleophilic substitution of N(6)-pivaloylated 9 with various alcohols in the presence of N-bromosuccinimide (NBS) gave the respective 4'-α-alkoxycordycepins 15a-21a as the major stereoisomers. Use of DAST in place of an alcohol gave the 4'-α-fluoro analogue 23a stereoselectively. Radical-mediated carbon-carbon bond construction was also applicable to 7, giving 4'-α-allylcordycepin (24a) and 4'-α-cyanoethylcordycepin (25) derivatives.  相似文献   

11.
The reaction of phenylmercuric chloride, 1-bromo-2,2,2-tetrafluoroethane and methanolic sodium methoxide in THF at ?35° gave PhHgCFBrCF3 in good yield. Thermolysis of this mercurial at 155° for 24 h in the presence of olefins gave gem-fluoro(trifluoromethyl)cyclopropanes, while its reaction with triethylsilane produced Et3SiCHFCF3.  相似文献   

12.
A dinuclear asymmetric zinc catalyst generated by mixing a 2:1 ratio of diethylzinc and 2,6-bis[5-2-diarylhydroxy methyl-1-pyrrolidinyl]-4-methylphenol has been contrasted with enzymes for the desymmetrization of some meso diols. The best ligand has a p-biphenylyl group as the aromatic substituent defining the chiral space. A series of 2-substituted propanediols were examined. The best acyl transfer agent proved to be vinyl benzoate. Diacylation normally did not occur. The phenyl substituted substrate gave 91-95% ee which compares favorably with the best ee of 92% reported for an enzymatic desymmetrization. The methyl substituted substrate gave significantly better results with the dinuclear zinc catalyst (89% yield, 82% ee) as compared to the best enzymatic esterification (70% yield, 60% ee). One case of a 1,4-diol, cis-1,2-bis(hydroxymethyl) cyclohexane, also gave much better results with the dinuclear zinc catalysts (93% yield, 91% ee) as compared to the reported enzymatic process (44% yield, 7% ee). A model to rationalize the results is presented.  相似文献   

13.
Friedel‐Crafts cycloalkylation of biphenyl with 2,3‐dichlorobutyltrichlorosilane(Cl3Si‐CH2CHClCHClCH3) at a temperature of 100°C in the presence of anhydrous aluminum chloride catalyst gave cyclized product, 9‐methyl‐9‐(2‐trichlorosilylethyl)fluorene 1a , in 33% yield. Methylation of 1a with nucleophilic reagent such as methyl magnesium chloride, gave 9‐methyl‐9‐(2‐trimethylsilylethyl)fluorene 2 while bromination of 2 with excess amount of bromine in DMF resulted in 2,7‐dibromo‐9‐methyl‐9‐(2‐trichlorosilyl)fluorene 3 in good yield. All the compounds were structurally identified by GC/MS, 1H and 13C‐NMR spectroscopy.  相似文献   

14.
己内酰胺三氟甲磺酸离子液体中甲苯的区域选择性硝化   总被引:1,自引:0,他引:1  
赵鑫  俞慧丽  葛继云 《应用化学》2009,26(8):905-908
以三氟甲磺酸和己内酰胺为原料制备了一种新型离子液体己内酰胺三氟甲磺酸([CP]OTf),用1HNMR、FTIR对其结构进行了表征。并把[CP]OTf作为催化剂和溶剂,用于催化甲苯的硝化反应。借助气相色谱研究了[CP]OTf的用量、硝化温度、反应时间、硝酸浓度和离子液体的重复使用等因素对甲苯硝化反应区域选择性和产率的影响。结果表明:在反应温度55℃,反应时间6h,甲苯与95%浓硝酸的摩尔比为1:1,离子液体用量占甲苯摩尔数的15%时,产物中邻对位硝基甲苯摩尔比为1.12,较浓硫酸催化硝化的相应值(1.67)显著降低,硝基甲苯的产率达到78.41%,[CP]OTf表现出了优良的催化活性和对位选择性。[CP]OTf) 重复使用5次,其催化活性及对位选择性变化很小。  相似文献   

15.
The 6-bromo ( 16 ), 6-fluoro ( 17 ), 7-bromo ( 14 ), and 7-fluoro ( 15 ) substituted 3-amino-3,4-dihydro-l-hydroxycarbostyrils were treated with concentrated hydrochloric and hydrobromic acids under reflux conditions. The 7-halogenated N-hydroxycarbostyrils ( 14,15 ) gave the normal rearrangement products, the 6,7-dihalolactams ( 18–21 ). The 6-halogenated compounds ( 16,17 ) yielded the corresponding 6,8-dihalolactams ( 22–24 ) under the same experimental conditions, with the exception of the hydrobromic acid reaction of the 6-fluoro derivative 17 which yielded a mixture of products. Based on the comparison of the nmr spectrum of the product mixture with those of two authentic compounds, the mixture was identified as consisting of the normal rearrangement product, the 8-bromo-6-fluorolactam ( 27 ) and the straightforward reduction product, the 6-fluorolactam ( 26 ) in a ratio of about 2:1. The latter compounds were prepared by an independent method of synthesis in which 2-amino-5-fluorophenylalanine ( 25 ) was acidified to yield the corresponding lactam 26 , followed by bromination to afford the 8-bromo-6-fluorolactam 27. A mechanism is proposed to interpret the experimental results of nucleophilic substitution with rearrangement and reduction which occur with the 6-fluoro compound 17 when exposed to bromide ions in strongly acidic solution.  相似文献   

16.
This article describes our first experiments for preparing dye‐labeled latex particles by the emulsion copolymerization of a 4/1 (w/w) mixture of vinyl acetate‐butylacrylate (VAc‐BA). We discuss the synthesis of acrylate derivatives of phenanthrene, anthracene, and pyrene [9‐acryloxymethyl phenanthrene ( 7 ), 9‐acryloxymethyl‐10‐methyl anthracene ( 8 ), and 1‐acryloxymethyl pyrene ( 10 )] and an allyl ether derivative of anthracene [9‐allyoxymethyl‐10‐methyl anthracene ( 9 )]. Although the phenanthrene derivative 7 gave latex particles with high monomer conversion and good dye incorporation, the pyrene acrylate and both anthracene comonomers strongly inhibited the free‐radical reaction. To assist our search for a dye that would serve as a useful energy acceptor for phenanthrene and without suppressing VAc‐BA polymerization, we also examined batch emulsion polymerization in the presence of a variety of dye derivatives—substituted anthracenes, acridines, a coumarin, and two benzophenone derivatives. All of the anthracene derivatives, as well as acridine, strongly inhibited monomer polymerization. The coumarin dye 7‐hydroxy‐4‐methyl coumarin ( 22 ) that had only limited solubility allowed more than 90% monomer conversion. Most promising were 2‐hydroxy‐5‐methyl benzophenone ( 23 ) and 4‐N,N‐dimethylamino benzophenone ( 24 ) that at 1 mol % in the monomer mixture permitted virtually quantitative monomer conversion to latex. 4′‐Dimethylamino‐2‐acryloxy‐5‐methyl benzophenone ( 25 ) copolymerized well with the VAc‐BA mixture, yielding latex particles in high yield and with a narrow size distribution. These dyes appear to be useful acceptor dyes for energy‐transfer experiments with phenanthrene. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 40: 1594–1607, 2002  相似文献   

17.
Regioselectivity occurs in the sulfur-bridging reactions of 2,2′-binaphthyl (1) and 1-(2-naphthyl)cyclohexene (7) by means of hydrogen sulfide and a chromia-alumina-magnesia catalyst (designated I) in a flow apparatus at 550°. Thus, 1 gives a higher yield (6.1%) of dinaphtho[1,2–6:2′,l'-d]thiophene from 1,1′-bridging than of dinaphtho[1,2-b:2′,3′-d]thiophene (3.4%) from l,3′-bridging. No product expected from 3,3′-bridging was identified. Substrate 7 undergoes both dehydrogenation and bridging to yield 2-phenylnaphthalene (8%), benzo[b]naphtho[2,1-d]thiophene (9%) from alpha bridging, and benzo[b]naphtho[2,3-d]thiophene (3%) from beta bridging into the naphthalene ring. Exploratory studies showed that either sulfided catalyst I or a sulfided molybdenum( VI ) oxide-alumina-cobalt( II ) oxide catalyst ( II ) effects hydrodehalogenation of various monohalo- and polyhaloarenes (where halo, X, is chloro or bromo) at 450–550°. In the biphenyl, phenanthrene, naphthalene, and pyrene systems, halogen was lost either under sulfur-bridging conditions or under hydrogenolysis conditions, i.e. with methanol as a reactant. For every substrate the parent arene was isolated or identified as a reaction product. In selected experiments, acid HX was also identified in the effluent. Use of hydrogen sulfide as a reactant led to formation of dibenzothiophene and phenanthro[4,5-bcd]thiophene as main products in the biphenyl and phenanthrene systems, respectively; while use of methanol as a reactant gave small amounts of methyl bromide (for X = Br) and methylarenes.  相似文献   

18.
A straightforward access to protected 8-fluoro nucleosides via metalation-electrophilic fluorination under heterogeneous reaction conditions is reported. This is the first synthesis of 8-fluoro-2'-deoxyribonucleoside derivatives. Phenylsulfonyl substituted nucleosides are accompanying byproducts, possibly indicating a competing radical process. Higher yields of 8-fluoro derivatives were obtained with 2'-deoxyribonucleosides, as compared to ribonucleosides. Deprotection of the hydroxyl groups leading to 8-fluoro-2'-deoxyadenosine using TASF in methylene chloride demonstrates the compatibility of desilylation with 8-fluoro substituted nucleosides. NMR data indicate a syn conformation of the 8-fluoro derivatives.  相似文献   

19.
A 6% slurry of polytetrafluoroethylene (PTFE) in 4% hydrofluoric acid and 1% nickel nitrate were used as modifiers for determination of selenium in soils by GF AAS. PTFE was used to remove silica from the soil sample, because this resulted in severe matrix effects. The temperature of fluorination, determined thermogravimetrically, was 600 °C. The yield of fluorination depends on the molar ratio of PTFE/silica, particle size, and the time and temperature of fluorination. The soil samples were pretreated with a small amount of concentrated hydrofluoric acid placed directly in the cup of autosampler. The results for the determination of selenium in the reference soil materials by means of the slurry-sampling technique and use of aqueous standards are in good agreement with the certified values. Received: 20 December 2000 / Revised: 3 April 2001 / Accepted: 5 April 2001  相似文献   

20.
A synthesis of 9-(2,3-dideoxy-2-fluoro-beta-D-threo-pentofuranosyl)adenine (1, FddA) via a 6-chloro-9-(3-deoxy-beta-D-erythro-pentofuranosyl)-9H-purine (9), which was readily obtained from inosine (5), is described. Fluorination at the C2'-beta position of the purine 3'-deoxynucleoside with diethylaminosulfur trifluoride was improved by the introduction of a 6-chloro group and proceeded in moderate yield. Purine 3'-deoxynucleoside derivatives were also subjected to nucleophilic reactions with triethylamine trihydrofluoride and gave the desired fluorinated nucleoside in good yield. The safety and yield of the fluorination process were greatly improved by the use of triethylamine trihydrofluoride. The influence of the sugar ring conformation and 6-chloro group on the rate of the nucleophilic reaction against elimination are also discussed.  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