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.
Eleven N-substituted carbazoles (CZ) with terminal vinyl groups were synthesized by five sequences of reaction: N-(p-vinylbenzyl)- and N-β-(vinyloxyethyl)CZ by N-alkylations of potassium CZ with corresponding chlorides; N-(β-acryloyloxyethyl)CZ by the esterification of N-(β-hydroxyethyl)CZ with acryloyl chloride; N-acrylamido-or methacryl-amido-methyl CZ from N-hydroxymethyl CZ and acryl- or methacrylamide; N-(3-acryloyl- or methacryloyl-oxy-2-hydroxypropyl)CZ and N-[3-(p- or m-vinylanilino)-2-hydroxypropyl] CZ from N-(2,3-epoxypropyl)CZ and acrylic or methacrylic acid and p- or m-vinylaniline, respectively; and 2-[β-(N-carbazyl)propionyloxy]ethyl acrylate or methacrylate by the Michael addition of CZ to 2-hydroxyethyl acrylate or methacrylate, followed by esterifications. The vinyl polymers with pendant carbazyl groups more or less distant from the polymer backbones, prepared by conventional radical or cationic polymerization procedures, indicated charge-transfer spectra with 2,4,7-trinitrofluorenone (TNF) in tetrahydrofuran (THF) solutions that are spread over most of the visible range.  相似文献   

2.
3-(2-Di-n-propylaminoethyl)pyrrole (1a) was prepared in good yield by reduction of pyrrole-3-(N,N-di-n-propylglyoxamide) (9) with lithium aluminum hydride. 3-(2-Di-n-propylaminoethyl)-5-acetylpyrrole (1b) was prepared by first acetylation of 1-p-toluenesulfonyl-3-(2-di-n-propylaminoethyl)pyrrole (6) followed by hydrolysis of the p-toluenesulfonyl substituent. The starting material 6 was prepared by homologation of 1-(p-toluenesulfonyl)pyrrole-3-carboxaldehyde (3) to the corresponding acetaldehyde followed by reductive amination of the latter.  相似文献   

3.
A novel heterotelechelic linear block copolymer of poly(ethylene glycol) (PEG) and poly(succinimide) (PSI) possessing both acetal and tert-butoxycarbonyl-amino (Boc-NH) terminals (Acetal-PEG-b-PSI-NH-Boc) with a narrow molecular weight distribution (MWD) was successfully prepared by the nucleophilic attack of triethylamine (TEA) to the poly(β-benzyl L-aspartate) (PBLA) segment of Acetal-PEG-b-PBLA-NH-Boc. Acetal-PEG-b-PBLA-NH-Boc with MWD of 1.07 was prepared by living anionic ring-opening polymerization of β-benzyl L-aspartate N-carboxy-anhydride with α-acetal-ω-amino PEG as a macroinitiator, followed by Boc protection. The subsequent conversion of PBLA segment to PSI was successfully carried out by reacting with the catalytic amount of TEA. The characterization by 1H NMR, GPC and IR demonstrates that the formation of poly(succinimide) proceeded completely without any remarkable side reactions. Acetal-PEG-b-PSI-NH-Boc thus obtained may have a potential utility as a targetable drug carrier in the field of drug delivery system.  相似文献   

4.
Some comments on the syntheses of 5-amino-m-xylene-2-sulfonic acid and 5-amino-m-xylene-4-sulfonic acid Treatment of 5-amino-m-xylene ( 1 ) with oleum led to a 55:45 mixture of 5-amino-m-xylene-2-sulfonic acid ( 2 ) and 5-amino-m-xylene-4-sulfonic acid ( 3 ). The structure of both isomers was proven by reaction of sulfur dioxide with the diazonium chlorides derived from 2-amino-5-nitro-m-xylene ( 5 ) and 4-amino-5-nitro-m-xylene ( 8 ) giving 5-nitro-m-xylene-2-sulfonyl chloride ( 6 ) and 5-nitro-m-xylene-4-sulfonyl chloride ( 9 ) respectively, followed by hydrolyses to the corresponding sulfonic acids 7 and 10 , and final Béchamp reductions. The sulfonic acid 2 was also prepared by sulfonation of 5-acetylamino-m-xylene ( 4 ) to 5-acetylamino-m-xylene-2-sulfonic acid ( 11 ) and subsequent hydrolysis. A further procedure for the synthesis of 3 was sulfonation of 5-amino-2-chloro-m-xylene ( 12 ) – prepared by Béchamp reduction of 2-chloro-5-nitro-m-xylene ( 13 ) – or of 5-amino-2-bromo-m-xylene ( 15 ) – prepared by bromination of 4 and subsequent hydrolysis – to 5-amino-2-chloro-m-xylene-4-sulfonic acid ( 16 ) and 5-amino-2-bromo-m-xylene-4-sulfonic acid ( 17 ) respectively, followed by hydrogenolysis.  相似文献   

5.
(s)-Pinanediol (1-methoxyvinyl)boronate ( 1 ) was prepared from (1-methoxyvinyl)-lithium and triisopropyl borate followed by (s)-pinanediol. Attempted reaction with (dichloromethyl)lithium failed, and reaction with butylmagnesium chloride followed by acetic acid yielded a mixture of diastereomers of (s)-pinanediol (1-methoxy-1-methyl-pentyl)boronate ( 2 ). (s)-Pinanediol (1-chlorovinyl)boronate ( 4 ) has been prepared by dehydrochlorination of (s)-pinanediol 1,1-dichloroethylboronate ( 3 ) with lithium chloride in dimethylformamide. Reaction of 4 with (dichloromethyl)lithium yielded (s)-pinanediol (1S)-(1,2-dichloroallyl)boronate ( 5 ) in 92% diastereomeric excess. Reaction of 5 with RMgX resulted in a 3 : 1 ratio of displacement of the 1-Cl from carbon by R to displacement of the entire 1,2-dichloroallyl group from boron by R. With lithium benzyl oxide, displacement of the 1-Cl from 5 failed entirely. Reaction of 4 with (dibromomethyl)lithium was inefficient and yielded a gross mixture of diastereomers.  相似文献   

6.
The conversion of 3-methyl-4-benzyl-4-chloro-2-pyrazolin-5-one 10b was catalyzed by a mixture of potassium fluoride and alumina to give syn-(methyl, benzyl)bimane 6 (62%) without detectable formation of the anti isomer, A6 [a 1 : 1 mixture (87%) of the isomers 6 and A6 was obtained when the catalyst was potassium carbonate]. In a similar reaction syn-(methyl,carboethoxymethyl)bimane 7 (15%) with the anti isomer A7 (36%) was obtained from 3-methyl-4-carboethoxymethyl-4-chloro-2-pyrazolin-5-one 10c . syn-(Methyl, β-acetoxyethyl)bimane 8 (70%) was obtained from 3-methyl-4-β-acetoxyethyl-4-chloro-2-pyrazolin-5-one 10d (potassium carbonate catalysis) and was converted by hydrolysis to syn-(methyl, β-hydroxyethyl)bimane 9 (40%). Acetyl nitrate (nitric acid in acetic anhydride) converted anti-(amino,hydrogen)bimane 11 to anti-(amino,nitro)bimane 15 (91%), anti-(methyl,hydrogen)bimane 13 to anti-(methyl,nitro)(methyl,hydrogen)bimane 16 (57%), and degraded syn-(methyl,hydrogen)bimane 12 to an intractable mixture. Treatment with trimethyl phosphite converted syn-(bromomethyl,methyl)bimane 17 to syn-(dimethoxyphosphinylmethyl,methyl)bimane 18 (78%) that was further converted to syn-(styryl,methyl)bimane 19 (29%) in a condensation reaction with benzaldehyde. Treatment with acryloyl chloride converted syn-(hydroxymethyl,methyl)bimane 20 to its acrylate ester 21 (22%). Stoichiometric bromination of syn-(methyl,methyl)bimane 1 gave a monobromo derivative that was converted in situ by treatment with potassium acetate to syn-(acetoxymethyl,methyl)(methyl,methyl)bimane 47 . N-Amino-μ-amino-syn-(methylene,methyl)bimane 24 (68%) was obtained from a reaction between the dibromide 17 and hydrazine. Derivatives of the hydrazine 24 included a perchlorate salt and a hydrazone 25 derived from acetone. Dehydrogenation of syn-(tetramethylene)bimane 26 by treatment with dichlorodicyanobenzoquinone (DDQ) gave syn-(benzo,tetramethylene)bimane 27 (58%) and syn-(benzo)bimane 28 (29%). Bromination of the bimane 26 gave a dibromide 29 (92%) that was also converted by treatment with DDQ to syn-(benzo)bimane 28 . Treatment with palladium (10%) on charcoal dehydrogenated 5, 6, 10, 11-tetrahydro-7H,9H-benz [6, 7] indazol [1, 2a]benz[g]indazol-7,9-dione 35 to syn-(α-naphtho)bimane 36 (71%). The bimane 35 was prepared from 1,2,3,4-tetrahydro-1-oxo-2-naphthoate 37 by stepwise treatment with hydrazine to give 1,2,4,5-tetrahydro-3H-benz[g]indazol-3-one 38 , followed by chlorine to give 3a-chloro-2,3a,4,5-tetrahydro-3H-benz[g]indazol-3-one 39 , and base. Dehydrogenation over palladium converted the indazolone 34 to 1H-benz[g] indazol-3-ol 36 . Helicity for the hexacyclic syn-(α-naphtho)bimane 36 was confirmed by an analysis based on molecular modeling. The relative efficiencies (RE) for laser activity in the spectral region 500–530 nm were obtained for 37 syn-bimanes by reference to coumarin 30 (RE 100): RE > 80 for syn-bimanes 3, 5, 18 , and μ-(dicarbomethoxy)methylene-syn-(methylene,methyl)bimane 22 : RE 20–80: for syn-bimanes 1,2,4,20,24,26 , and μ-thia-syn-(methylene,methyl)bimane 50 : and RE 0-20 for 26 syn-bimanes. The bimane dyes tended to be more photostable and more water-soluble than coumarin 30. The diphosphonate 18 in dioxane showed laser activity at 438 nm and in water at 514 nm. Presumably helicity, that was demonstrated by molecular modeling, brought about a low fluorescence intensity for syn-(α-naphtho)bimane 36 , Φ0.1, considerably lower than obtained for syn-(benzo)bimane 28 , Φ0.9.  相似文献   

7.
4-Nitro-N-vinylphthalimide ( 4 ) was synthesized by two different procedures. Compound 4 was not polymerizable or copolymerizable by AIBN. Poly(N-vinylphthalimide) ( 17 ) was prepared and partially nitrated at 10–25°C. N,N′-(1,2-Ethanediyl)bis(4-nitrophthalimide) ( 15 ) and N,N′-(1,3-propanediyl)bis(4-nitrophthalimide) ( 16 ) were prepared by the condensation of the corresponding diamine with phthalic anhydride followed by nitration of the condensation products. 4-Nitrophthalic anhydride was prepared by the hydrolysis of 15 . Four styrene-substituted phthalimide monomers were synthesized. These include N-(4-vinylphenyl)phthalimide ( 25a ), N-(4-vinylphenyl)-3-fluorophthalimide ( 25b ), N-(4-vinylphenyl)-3-nitrophthalimide ( 25c ), and N-(4-vinylphenyl)-4-nitrophthalimide ( 25d ). Monomers 25a and 25b were polymerized by freeradical initiator (AIBN), whereas monomers 25c and 25d were not polymerizable or copolymerizable by AIBN due to a strong inhibitive effect exerted by the nitrophthalimide group. Monomers 25c and 25d were cationically polymerized (BF3·OEt2). Monomer 25b and styrene were copolymerized and their reactivity ratios were r1 = 1.7 and r2 = 0.55, respectively. The prepared polymers are useful as backbone polymers for grafting living anionic polymers.  相似文献   

8.
Usually, ortho lithiation of Kagan's template 1 and quenching with electrophiles leads highly diastereoselectively to planar‐chiral 1,2‐disubstituted ferrocenes. Surprisingly, lithiation of 1 with lithium diisopropylamide (LDA) followed by addition of paraformaldehyde afforded regioisomer (+)‐{[S(S)][4‐(2‐hydroxyethyl)phenyl]sulfinyl}ferrocene ( 2 ), which was converted to (+)‐{[S(S)]{4‐{2‐[(methylsulfonyl)oxy]ethyl}phenyl}sulfinyl}ferrocene ( 3 ) (Scheme 1). The desired diastereoisomer (l)‐1‐(hydroxymethyl)‐2‐(p‐tolylsulfinyl)ferrocene ( 5 ) in turn could also be obtained by ortho lithiation of 1 with LDA but by quenching with DMF to yield aldehyde 4 first, which then was reduced with NaBH4 to 5 . Finally, target compound (l)‐1‐[(dimethylamino)methyl]‐2‐(p‐tolylsulfinyl)ferrocene ( 6 ) was obtained by substitution of the OH group of 5 under mild conditions or directly by ortho lithiation of 1 with lithio‐2,4,6‐triisopropylbenzene (=2,4,6‐triisopropylphenyl)lithium; LTP) followed by quenching with N,N‐dimethylmethyleneiminium chloride. At low temperatures, reaction of 1 with LDA leads, via the preferred diastereoisomeric transition state ‘exo’‐ 7 and under extrusion of a (diisopropylamine)lithium complex of type 8 , in a highly selective manner, to diastereoisomeric ortho‐lithiated chelate (l)‐ 9 (Scheme 2). The reaction of 1 to 2 is explained by a rearrangement of (l)‐ 9 to {[S(S)] [4‐(lithiomethyl)phenyl]sulfinyl}ferrocene 10 , which is acid‐catalyzed by coordinated diisopropylamine in complexes of type 8 . This rearrangement is not observed if LTP is used as base or, in case LDA is applied, if the electrophile is sufficiently reactive at low temperatures.  相似文献   

9.
2-Acetamido-1,3,4,6-tetra-O-acetyl-2-deoxy-α-D-glucopyranose (I), and its analogs specifically mono (trideuterioacetylated) at O-1 (III), at N-2 (II), at O-4 (IV) and at O-6 (V), have been examined by high-resolution mass spectrometry. From the elemental compositions of the fragment ions, the mass-number shifts resulting from deuterium incorporation and analysis of metastable transitions, it has been possible to specify in detail the fragmentation pathways undergone by this molecule. The principal degradations of I proceed by initial rapid decomposition of the molecular ion (whose intensity is insignificant) by three routes: (i) by loss of the C-1 acetoxyl group as a radical to give the glycosyl cation (a), (ii) by loss of the 1-acetyl group as a radical to give an acyclic ion m/e 346 (b) and (iii) by loss of a C-6 fragment and acetic acid derived from the 3-acetate group to give m/e 241 (c).  相似文献   

10.
(R)—(+)-Limonene was photooxidized in the presence of Rose Bengal as catalyst. After TLC isolation, the hydroperoxides formed were separated directly by HRGC and analyzed by MS (El; Cl). Each hydroperoxide isomer was then isolated by HPLC for structure determination which after reduction of the HOO group with sodium borohydride was performed by 1H-NMR and 13C-NMR. Six hydroperoxide isomers formed by oxidation of the endocyclic double bond were identified. The compounds eluted from the HRGC column in the following order (proportions are given in brackets) I (40.1%) (1S, 4R)-p-mentha-2, 8-diene 1-hydroperoxide; II (5.8%) (1R, 4R)-p-mentha-2, 8-diene 1-hydroperoxide; III (20.6%) (2R, 4R)-p-mentha-[1(7), 8]-diene 2-hydroperoxide; IV (8.5%) (2R, 4R)-p-mentha-6, 8-diene 2-hydroperoxide; V (4%) (2S, 4R)-p-mentha-6, 8-diene 2-hydroperoxide; and VI (21.0%) (2S, 4R)-p-mentha-[1(7), 8]-diene 2-hydroperoxide. Direct HRGC separation of the limonene hydroperoxides offers, inter alia, the possibility of determining their flavor qualities by HRGC/effluent sniffing.  相似文献   

11.
2-(9H-Xanthen-9-ylmethyl)-1H-benzimidazole ( 2a ) was prepared by condensing 9H-xanthene-9-acetic acid ( 1a ) with 1,2-benzenediamine. Similarly, 2-(9H-thioxanthen-9-ylmethyl)-1H-benzimidazole ( 2b ) and its S,S-dioxide ( 2d ) were obtained. Compound 2d was also prepared by oxidizing 2b with hydrogen peroxide in acetic acid. Heating of 9H-thioxanthene-9-acetic acid 10-oxide ( 1c ) with 1,2-benzenediamine gave 9-methylene-9H-thioxanthene ( 3 ). 2-(9H-Thioxanthen-9-ylmethyl)-1H-benzimidazole S-oxide ( 2c ) was obtained by oxidizing 2b with m-chloroperbenzoic acid in acetone.  相似文献   

12.
A pressure‐controlled procedure for the SN1 reaction of rac‐1‐[(dimethylamino)methyl]‐2‐(tributylstannyl)ferrocene ( 1 ) to rac‐1‐(phthalimidomethyl)‐2‐(tributylstannyl)ferrocene ( 2 ) was developed. Pd0‐Catalyzed Stille coupling of 2 with iodobenzene afforded rac‐1‐phenyl‐2‐(N‐phthalimidomethyl)ferrocene ( 5 ) in 74% yield; after trace enrichment by crystallization of the combined mother liquors, one single crystal of each, 5 , catalysis intermediate trans‐iodo(σ‐phenyl)bis(triphenylarsino)palladium(II) ( 7 ), trans‐diiodobis(triphenylarsino)palladium(II) ( 8 ), and rac‐2,2′‐bis(phthalimidomethyl)‐1,1′‐biferrocene ( 9 ) could be isolated by crystal sorting under a microscope and characterized by X‐ray crystal structure analysis. Furthermore, 5 was deprotected to amine ( 11 ), which does even survive the Birch reduction to rac‐1‐(aminomethyl)‐2‐(cyclohexa‐2,5‐dienyl)ferrocene ( 12 ).  相似文献   

13.
The cob (I)alamin- ( 1(I) ) and the heptamethyl cob(I)ynnate- ( 2(I) ) catalyzed transformation of an epoxide to the corresponding saturated hydrocarbon 3→4→5 is examined (see Schemes 1 and 3–5). Under the reaction conditions, the epoxyalkyl acetate 3 is opened by the catalysts with formation of appropriate (b?-hydroxyalkyl)-corrinoid derivatives ( 13 , 14 , 17 , 18 , see Schemes 12 and 14). Triggered by a transfer of electrons to the Co-corrin-π system, the Co, C-bond of the intermediates is broken, generating the alkenyl acetate 4 (cf. Schemes 12 and 14) following an electrofugal fragmentation (cf. Schemes 2 and 12). The double bond of 4 is also attacked by the catalysts, leading to the corresponding alkylcorrinoids ( 15 , 19 , see Schemes 12 and 14) which in turn are reduced by electrons from metallic zinc, the electron source in the system, inducing a reductive cleavage of the Co, C-bond with production of the saturated monoacetate 5 (see Schemes 2, 5 and 12). In the cascade of steps involved, the transfer of electrons to the intermediate alkylcorrinoids ( 13–15 , 17–19 , see Schemes 12 and 14) is shown to be rate-limiting. Comparing the two catalytic species 1(I) and 2(I) , it is shown that the ribonucleotide loop protects intermediate alkylcobalamins to some extent from an attack by electrons. The protective function of the ribonucleotide side-chain is shown to be present in alkylcobalamins existing in the base-on form (cf. Chap. 4 and see Scheme 14).  相似文献   

14.
Three N‐Boc‐protected amino acids substituted with a mesityl (=2,4,6‐trimethylphenyl) group were synthesized in enantiomerically pure form, either by asymmetric epoxidation or by aminohydroxylation as the source of chirality. The 3‐mesityloxirane‐2‐methanol 7 , easily available in high enantiomer purity by Sharpless epoxidation, was converted into 3‐{[(tert‐butoxy)carbonyl]amino}‐3‐mesitylpropane‐1,2‐diol 9 by a regio‐ and stereoselective ring opening with an ammonia equivalent (sodium azide or benzhydrylamine), followed by hydrogenation and in situ treatment with (Boc)2O (Boc=[(tert‐butoxy)carbonyl]) (Scheme 3). Oxidative cleavage of the diol fragment in 9 afforded N‐[(tert‐butoxy)carbonyl]‐α‐mesitylglycine 1 of >99% ee. This amino acid was also prepared in enantiomerically pure form starting from 2,4,6‐trimethylstyrene ( 11 ) by a regioselective Sharpless asymmetric aminohydroxylation, followed by a 2,2,6,6‐tetramethylpiperidin‐1‐yloxyl (TEMPO)‐catalyzed oxidation (Scheme 4). On the other hand, 1‐[(tert‐butoxy)carbonyl]‐2‐{{[(tert‐butyl)dimethylsilyl]oxy}methyl}‐3‐mesitylaziridine 14 was prepared from 9 by a sequence involving selective protection of the primary alcohol (as a silyl ether), activation of the secondary alcohol as a mesylate, and base‐induced (NaH) cyclization (Scheme 5). The reductive cleavage of the aziridine ring (H2, Pd/C), followed by alcohol deprotection (Bu4NF/THF) and oxidation (pyridinium dichromate (PDC)/DMF or (TEMPO)/NaClO) provided, in high yield and enantiomeric purity, N‐[(tert‐butoxy)carbonyl]‐β‐mesitylalanine 2 . Alternatively, the regioselective ring opening of the aziridine ring of 14 with lithium dimethylcuprate, followed by silyl‐ether cleavage and oxidation lead to N‐[(tert‐butoxy)carbonyl]‐β‐mesityl‐β‐methylalanine 3 . A conformational study of the methyl esters of the N‐Boc‐protected amino acids 1 and 3 carried out by variable‐temperature 1H‐NMR and semi‐empirical (AM1) calculations shows the strong rotational restriction imposed by the mesityl group.  相似文献   

15.
M. Hinge 《Colloid Journal》2007,69(3):342-347
The aim of the performed work is to produce anionic core-shell poly(styrene-co-N-isopropylacrylamide) colloids with an N-isopropylacrylamide (NIPAM) content in the range from 5 to 30 mol %. Different batches of poly(styrene-co-NIPAM) colloids (poly(ST-co-NIPAM) colloids) are produced with varying NIPAM mol %, and the produced poly(ST-co-NIPAM) colloids are characterized by dynamic light scattering and scanning electron microscopy. Results show that the produced colloids have a core-shell morphology with a poly(styrene) core and a poly(NIPAM) shell. The swelling ratio of the produced poly(ST-co-NIPAM) colloids is similar to the swelling ratio found for similar poly(ST-co-NIPAM) colloids produced by the two-step seeded polymerization process. The text was submitted by the author in English.  相似文献   

16.
A series of perfluorocyclobutyl (PFCB) aryl ether‐based amphiphilic diblock copolymers containing hydrophilic poly(acrylic acid) (PAA) and fluorophilic poly(p‐(2‐(p‐tolyloxy)perfluorocyclobutoxy)phenyl methacrylate) segments were synthesized via successive atom transfer radical polymerization (ATRP). 2‐MBP‐initiated and CuBr/N,N,N,N,N″‐pentamethyldiethylenetriamine‐catalyzed ATRP homopolymerization of the PFCB‐containing methacrylate monomer, p‐(2‐(p‐tolyloxy)perfluorocyclobutoxy)phenyl methacrylate, can be performed in a controlled mode as confirmed by the fact that the number‐average molecular weights (Mn) increased linearly with the conversions of the monomer while the polydispersity indices kept below 1.38. The block copolymers with narrow molecular weight distributions (Mw/Mn ≤ 1.36) were synthesized by ATRP using Br‐end‐functionalized poly(tert‐butyl acrylate) (PtBA) as macroinitiator followed by the acidolysis of hydrophobic PtBA block into hydrophilic PAA segment. The critical micelle concentrations of the amphiphilic diblock copolymers in different surroundings were determined by fluorescence spectroscopy using N‐phenyl‐1‐naphthylamine as probe. The morphology and size of the micelles were investigated by transmission electron microscopy and dynamic laser light scattering, respectively. © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2010  相似文献   

17.
A number of 2,4-disubstituted pyrrolo[3,2-d]pyrimidine N-5 nucleosides were prepared by the direct glycosylation of the sodium salt of 2,4-dichloro-5H-pyrrolo[3,2-d]pyrimidine (3) using 1-chloro-2-deoxy-3,5-di-O-(p-toluoyl)-α-D -erythropentofuranose (1) and 1-chloro-2,3,5-tri-O-benzyl-α-D-arabinofuranose (11) . The resulting N-5 glycosides, 2,4-dichloro-5-(2-deoxy-3,5-di-O-(p-toluoyl) -β-D-erythropentofuranosyl)-5H-pyrrolo-[3,2-d]pyrimidine (4) and 2,4-dichloro-5-(2,3,5-tri-O-benzyl-β-D-arabinofuranosyl-5H -pyrrolo [3,2-d)pyrimidine (12) , served as versatile key intermediates from which the N-7 glycosyl analogs of the naturally occurring purine nucleosides adenosine, inosine and guanosine were synthesized. Thus, treatment of 4 with methanolic ammonia followed by dehalogenation provided the adenosine analog, 4-amino-5-(2-deoxyerythropentofuranosyl) -5H-pyrrolo[3,2-d]pyrimidine (6) . Reaction of 4 with sodium hydroxide followed by dehalogenation afforded the inosine analog, 5-(2-deoxy-β-D-erythropentofuranosyl) -5H-pyrrolo[3,2-d]pyrimidin-4(3H)-one (9) . Treatment of 4 with sodium hydroxide followed by methanolic ammonia gave the guanosine analog, 2-amino-5-(2-deoxy-β-D-erythropentofuranosyl) -5H-pyrrolo[3,2-d]pyrimidin-4(3H)-one (10) . The preparation of the same analogs in the β-D-arabinonucleoside series was achieved by the same general procedures as those employed for the corresponding 2′-deoxy-β-D-ribonucleoside analogs except that, in all but one case, debenzylation of the sugar protecting groups was accomplished with cyclohexene-palladium hydroxide on carbon, providing 4-amino-5-β-D-arabinofuranosyl-5H-pyrrolo [3,2-d]pyrimidin-4(3H)-one (18) . Structural characterization of the 2′-deoxyribonucleoside analogs was based on uv and proton nmr while that of the arabinonucleosides was confirmed by single-crystal X-ray analysis of 15a . The stereospecific attachment of the 2-deoxy-β-D-ribofuranosyl and β-D-arabinofuranosyl moieties appears to be due to a Walden inversion at the C1 carbon by the anionic heterocyclic nitrogen (SN2 mechanism).  相似文献   

18.
The rate of hydrolysis of N-methyl-N-nitrosoamides by aqueous alkalis varies greatly. Methylnitrosourea (1) is hydrolyzed rapidly by aqueous KOH-solutions at low temperatures to give a high yield of diazomethane. Under similar conditions, N,N′-dimethyl-N,N′-dinitroso-oxamide (3) is hydrolyzed more slowly, but also gives a good yield of diazomethane. N,N′-Dimethyl-N,N′-dinitrosoterephthal-amide (4) , and (N-methyl-N-nitroso)-4-amino-4-methyl-2-pentanone (5) are less easily hydrolyzed by aqueous KOH-solutions. N-Methyl-N-nitroso-p-toluenesulfonamide (2) was the least reactive out of those tested. The hydrolysis of diazomethane in toluene with aqueous bases follows first order kinetics. The hydrolysis rate is greatly influenced by the concentration and strength of the base and temperature.  相似文献   

19.
9H-Dithieno[2,1 -b:4,5-c′]tropylium ion (III) and 4ii-dithieno[1,2-b:4,5-c′]tropylium ion (IV) have been synthesized by ring-closure of 1-(4-carboxy-3-thienyl)-2-(3′-thienyl)ethane (IX) and 1-(4-carboxy-3-thienyl)-2-(2′-thienyl)ethane (XVI), respectively, followed by bromination-debrom-ination to 9H-cyclohepta[2,1-b:4,5-c′] dithiophen-9-one (XI) and 4H-cyclohepta[1,2-b:4,5-c′]-dithiophen-4-one (XVIII), and finally by reduction and hydride transfer. The tropylium ions III and IV were less stable than the [b,b′]-fused isomers previously studied.  相似文献   

20.
Elastic response of the disordered phase between crystal lamellae in semicrystalline polymers is modelled on the assumption that the stress is transferred by bridging (tie) molecules. The deformation characteristics of short poly(methylene) (PM) bridges were computed by using two methods: (a) the single‐molecule loading by molecular mechanics (MM) calculations and (b) the chain‐ensemble averaging by lattice simulations. The energy elastic functions ensuing from both methods differ considerably. In MM the loading of chains containing numerous gauche defects by an external force F yields the sawtooth‐like profile of the force (F)–length (R) functions brought about by the stress‐induced gauchetrans conformational transitions. The Young's moduli E of PM chains containing several gauche defects can be less than 1% of the all‐trans value ET; by elimination of the defects the moduli steeply increase. In contrast, the ensemble‐averaging approach gives a smooth increase of the (positive) elastic force f with chain length R and a decrease of the (negative) energy component of the elastic force fU with R. Both energy deformation mechanisms, single‐chain loading (by F) and statistical (by fU), are complementary and can simultaneously be operative in the interlamellar (IL) phase. Their proportion in the stretching process should depend on the chain mobility and structural homogeneity (history) of the sample, particularly on the presence of the so‐called rigid amorphous fraction in the IL phase.  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