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.
α-(Aminornethylene)-9H-purine-6-acetamide ( 3a ) and the corresponding ethyl acetate 9 have been synthesized by catalytic hydrogenation of 6-cyanomethylenepurine derivatives 2 and 7 which were obtained by the substitution of 6-chloropurine derivatives with α-cyanoacetamide and ethyl cyanoacetate, respectively. Substitution of α-(aminomethylene)-9-(tetrahydrofuran)-9H-purine-6-acetamide ( 3b ) with amines gave the corresponding N-alkyl- and N-arylamines 5 , which were treated with acid to give N-substituted α-(aminomethylene)-9H-purine-6-acetamides 6 . Substitution of 9 with amines gave the corresponding N-alkyl- and N-aryl substituted amines 10 .  相似文献   

2.
Azatryptophane homologues, 4-(imidazo[1,2-a]pyridinyl-3)- 9a-9f and 4-(imidazo[1,2-a]pyrimidinyl-3)-4-oxohomoalanine derivatives 9g-91 , were prepared from N,N-dimethyl-N′-(pyridinyl-2)- 6a-6f and N,N-dimethyl-N-(pyriniidinyl-2)formamidines 6g-6i , and (S)-N-trifluoroacetyl-5-bromo-4-oxonorvaline methyl ester ( 2 ) and its (R,S)-isomer.  相似文献   

3.
Mono- and bisthioureas were synthesized based on N-(9-anthrylmethyl)-substituted alkylamines and diamines. Their luminescent and complexing properties were studied by the electronic, IR, and NMR 1H spectroscopy. N-(9-Anthrylmethyl)-N-(6-{9-anthrylmethyl[(phenylamino)carbothioyl]amino}hexyl)-N′-phenylthiourea was shown to be a highly effective fluorescent chemosensor for Hg2+ cations.  相似文献   

4.
The synthesis of N,N-diethyl-9-carbazolylacetamide ( 6 ), (R,S)- and (R)-N,N-diethyl-2-(9-carbazolyl)propanamide ( 7 ), N,N′-dimethyl-N,N′-di-(9-carbazolylacetyl)-1,2-ethanediamine ( 11 ), and (R)-N,N′-dimethyl-N,N′-di[2-(9-carbazolyl)propanoyl]-1,2-ethanediamine ( 13 ) is reported. The racemic compound, (R,S)-2-(9-carbazolyl)propanoic acid ( 2 ), was resolved by partial crystallization of the diastereomeric salts formed between 2 and (+)-α-methylbenzylamine. The 1H-NMR spectra of 6 and 7 showed magnetic nonequivalence of the chemically equivalent protons of the methyl and methylene groups in 6 and 7 due to partial double bond character of the amide bond. The upfield resonances corresponding to the two sets of methyl and methylene protons were assigned by the aromatic solvent-induced shift (ASIS) method to the protons anti to the carbonyl oxygen in the conformation of amide bond in 6 and 7 . The 1H-NMR spectra of 11 and (R)- 13 were used to determine the population of anti-anti, anti-syn (syn-anti) and syn-syn conformers in the structures of these dimer model compounds; the relative conformer populations were 0.45:0.47:0.08 and 0.28±0.02:0.29±0.01:0.43±0.01 in 11 and (R)- 13 .  相似文献   

5.
Synthesis of N-Methyl- and N,N-Dimethylmerucathine and of N-Methyl- and N,N--Dimethylpseudomerucathine Starting from L -Alanine Starting form L -alanine, N-methylmerucathine (= (3R,4S)-4-(methylamino)1-phenyl-1-penten-3-ol; (3R,4S,)- 6 ), N,N-dimethylmerucathine (= (3R,4S)-4-(dimethylamino)-1-phenyl-1-penten-3-ol; (3R,4S)- 9 ), N-methylpseudomerucathine (= (3S,4S)-4-(methylamino)-1-phenyl-1-penten-3-01; (3S,4S)-6), and N,N-dimethylpseudomerucathine (= (3S,4S)-4-(dimethylamino)-1-phenyl-1-penten-3-ol; (3S,4S)- 9 ) were synthesized. The four compounds were analyzed by HPLC and compared with a natural khat extract.  相似文献   

6.
A novel tetradentate amine ligand namely N,N,N′,N″,N‴;,N‴;‐hexaoligo(ethylene glycol) triethylenetetramine (HOEGTETA) was employed in the homogenous ATRP of MMA in anisole using CuBr and CuBr2 as the catalyst and ethyl 2‐bromoisobutyrate (EBiB) as an initiator. The effect of the polymerization temperature and the various ratios of Cu(I) to Cu(II) were investigated in detail. Moreover, we demonstrated the ATRP of MMA by using only Cu(II) in the absence of any free radical initiator, reducing agent, or air. The ATRP of MMA with the use of only Cu(II) and HOEGTETA or N,N,N′,N″,N″‐pentamethyldiethylenetriamine (PMDETA) resulted in well‐defined PMMA.

  相似文献   


7.
6-Methyl-6,7,8,9-tetrahydro-4H-pyrido[1,2-a]pyrimidin-4-ones 1-5 were subjected to Vilsmeier-Haack acylation with complexes of phosphoryl chloride and different amides. Acylation at position 9 of the pyridopyrim-idines was successful with the iminium salt formed in situ from N-formylpiperidine, N-methylformanilide or N,N-diethylbenzamide, but unsuccessful with the iminium salt formed from N,N-diethylacetamide or N,N-di-ethylisobutyramide, respectively. The iminium salt formed from formanilide, N-methylpyrrolidinone or formamide reacted only with those tetrahydropyridopyrimidinones which contain a strongly electronegative substituent (e.g. CN or CO2Et) in position 3. With the latter derivatives, the 9-phenylaminomethylene group could be introduced using N,N-diphenylformamide or in a “one-pot” procedure with aniline and triethyl orthoformate. Ethanolysis of 9-N-methyl-N-phenylaminomethylene derivatives 15 and 19 afforded 9-ethoxy-methylene compounds 26 and 27 in the presence of hydrogen chloride. The structures of the 9-substituted 6-methyltetrahydropyridopyrimidin-4-ones 14-25 were elucidated by means of uv, 1H and 13C nmr spectroscopy. 9-Piperidinomethylene 14 , 9-(N-methyl-N-phenylaminomethylene 15-19 and 9-(N-methyl-2-pyrrolidinylidene) 21 derivatives exist as E geometric isomers. 9-Phenylaminomethylene-6-methyl-4-oxo-6,7,8,9-tetra-hydro-4H-pyrido[1,2-a]pyrimidine-3-carbonitrile 20 displays a solvent-dependent E-Z isomerism. The bis-compound 25 contains both E and Z geometric exo C ? CH double bonds. 9-Benzoyl derivatives 23 and 24 exist predominantly as the 1,6,7,8-tetrahydropyridopyrimidin-4-one tautomer.  相似文献   

8.
Nucleophilic attack by N2 of N1 N2-diarylformamides 1a-c on C-2 of chloranil (2) and subsequently by N1 on C-1 of 2 initiates the formation of benzimidazolinones 8a-c. In contrast, when 1b-e is reacted with 2,3-dichloro-1,4-naphthoquinone (9) , both chlorine atoms are successively substituted by the two nitrogen atoms and 2-(arylamino)-3-(N-formylarylamino)-1,4-naphthoquinones 13b-e result, which (probably via their cyclic tautomers 12b-e ) may be cyclodehydrogenated to form N1,N3-diarylnaphtho[2,3-d] imidazoline-2,4,9-triones (as 14b,c ). On the other hand, N1,N2diarylacetamidines 15a-d attack 2 and 9 at C-2 with N2 but subsequently exert nucleophilic character at the acetamidine α-carbon attacking C-1 of 2 and 9 , respectively, thus forming 1-aryl-2-(arylimino)-3a-hydroxy-2,3,3a,6-tetrahydro-1N-indol-6-ones 18a-d and 3-aryl-2-(arylimino)-9b-hydroxy-2,3,5,9b-tetrahydro-1-H-benz[e]indol-5-ones 19b,c , respectively. The latter may be thermally dehydrated to the fully conjugated 2,5-dihydro-3H-benz[e]indol-5-ones 20b,c. Unambiguous structural assignments for 18b and 20c are made on the basis of X-ray crystal structure analyses.  相似文献   

9.
Abstract

cis- and trans-9-Phenylselenoxanthene-N-(arylsulfonyl)selenilimines were synthesized and isolated. Their stereochemistry was ascertained from the NMR spectra. Cis isomers reacted with chloramine-T or -B by an SN 2 type substitution to form trans isomers, but the reverse reaction did not take place. When trans isomers were refluxed in toluene they underwent intermolecular 1,4 rearrangement to give 9-arylsulfonamido-9-phenylselenoxanthene. The cis isomers neither rearranged nor isomerized. On treatment with DABCO, both isomers rearranged intermolecularly to 9-(N-arylsulfonamido)selenoxanthenes at room temperature. Hydrolysis of both isomers yielded trans-9-phenylselenoxanthene 10-oxide. Reactions with p-methoxyphenylmagnesium bromide or methylmagnesium iodide afforded 9-(p-methoxyphenyl)-9-phenylselenoxanthene or 9-phenylselenoxanthene as a main product, respectively.  相似文献   

10.
The reaction of 6-chloro-2-[1-methyl-2-(Mmemylthiocarbamoyl)hydrazino]quinoxaline 4-oxide 5 with acetic anhydride or trifluoroacetic anhydride resulted in dehydrative cyclization to give 2-(N-acetyl)-memylamino-8-chloro-4-methyl-4H-1,3,4-thiadiazino[5,6-b]quinoxaline 6 or 8-chloro-2-(N-trifluoroacetyl)methylamino-4-methyl-4H-1,3,4-thiadiazino[5,6-b]quinoxaline 9 , respectively. The oxidation of compound 6 or 9 with 2-fold molar amount of m-chloroperbenzoic acid afforded the 4H-1,3,4-thiadiazino-[5,6-b]quinoxaline 1,1-dioxide 8 or 13 , respectively. The acetyl group of compound 6 was hardly hydrolyzed, but the trifluoroacetyl group of compound 9 was easily hydrolyzed to change into 8-chloro-4-methyl-2-memylamino-4H-1,3,4-thiadiazino[5,6-b]quinoxaline 10 . The acylation of compound 10 with acetic anhydride, trifluoroacetic anhydride, phenyl isocyanate, and chloroacetyl chloride furnished the 2-(N-acetyl)methylamino 6 , 2-(N-trifluoroacetyl)methylamino 9 , 2-(1-methyl-3-phenylureido) 11 , and 2-(N-chloroacetyl)methylamino 12 derivatives, respectively.  相似文献   

11.
The reactions of phenanthrene 9,10-imine ( 1 ) with aromatic aldehydes, benzoic acids and acetylenedi-carboxylic esters were investigated. The aldehydes were shown to give 1-[N-(arylmethylidene)-9-phenanthreneamine-10-yl]-1a,9b-dihydrophenanthro[9,10-b]azirine 2. The ‘dimeric’ structure of these products was established by X-ray diffraction analysis. The carboxylic acids proved to form in the presence of dicyclohexylcarbodiimide, N-aroylphenanthrene 9,10-imines 7 , that readily undergo rearrangement to N-aroyl-9-phenanthrenamines 8. Esters of acetylenedicarboxylic acid gave the corresponding esters of (Z)-2-(1a,9b-dihydrophenanthro[9,10-b]azirine-1-yl)-2-butendioic acid 10 .  相似文献   

12.
The hypothesis which explains the diastereoselectivity of the 1,3-dipolar cycloaddition of the N-glycosylnitrones 1 – 3 leading to the 5,5-disubstituted isoxazolidines 4 – 6 on the basis of a kinetic anomeric effect predicts that nucleophiles should add to N-glycosylnitrones with a high degree of diastereoselectivity. To test this prediction, the nucleophilic addition of lithium and potassium dialkylphosphites to the crystalline (Z)-nitrone 11 , prepared from oxime 9 and (benzyloxy)acetaldehyde has been examined. The addition of lithium phosphites gave the N-glycosyl-N-hydroxyaminophosphonates 12 – 16 (d. e. 78–92%) in high yields (Scheme 4). The addition of potassium phosphites showed a much lower diastereoselectivity. Glycoside cleavage, hydrogenolysis, and dealkylation of 12 – 16 gave (+)-(S)-phosphoserine (+)- 19 (34–45% from 9 ). Its absolute configuration was confirmed by an X-ray analysis of the N-(3,3,3-trifluoro-2-methoxy-2-phenylpropionyl) derivative 24 . Similarly, the crystalline nitrone 25 gave the N-glycosyl-N-hydroxyaminophosphonate 26 , which was transformed into (+)-(S)-phosphovaline (+)- 31 (42% from 9 ). The diastereoselectivity of the nucleophilic addition and the enantiomeric purity of (+)- 31 were determined by the analysis of the derivative 30 (d.e. 92%) and 32 (d.e. 93%), respectively. The addition of lithium diethyl phosphite to the nitrone 33 , prepared in situ, gave the N-glycosyl-N-hydroxyaminophosphonate 34 , (41%; d.e. 91%), which was transformed in (+)-(S)-phosphoalanine (+)- 37 (21% from 9 ).  相似文献   

13.
Structural correlations have been carried out from 13C chemical shifts (δ) and by analysis of 1J(CH) coupling constants, and the conformation about the glycosidic bond has been studied by means of the 3J(CH) vicinal coupling constants between C-8 and H-1′ of some adenine nucleosides such as adenosine (Ado), N(7)-β-D-ribofuranosyladenine (N(7)-Ado), N(9)- and N(7)-β-D-xylofuranosyladenine (N(9)-xylAde and N(7)-xylAde), N(9)-(3-chloro-3-deoxy-β-D-xylofuranosyl)adenine (3′-Cl-xylAde) and N(9)-(2-chloro-2-deoxy-β-D-arabinofuranosyl)adenine (2′-Cl-araAde). The analysis of the influence on δ13C of the nature and configuration of the substituent in the carbohydrate fragment of the molecule has revealed two types of effects, namely, 1,2-cis and 1,2-trans. This approach, as well as the 3J(CH) values and the analysis of the C-3′-endo?C-2′-endo equilibrium of the carbohydrate fragment of nucleosides, and circular dichroism (CD) data, provides important information on the conformation about the glycosidic bond. The magnitudes of 3J(C-4, H) are indicative of the position of attachment of the carbohydrate fragment to the heterocyclic base.  相似文献   

14.
Inclusion of an N-6 acetate into the ambident nucleophile, sodium 1-N-methylisoguanide, resulted in a shift in alkylation preference from N-9 to N-3. The preparation of the N-6 acetate of 1-N-methylisoguanine, 9-N-acetyl-1-N-methylisoguanine (4) , and related acetate analogs are described and evidence presented for their structural determination. Analysis of long range 13C-1H coupling data facilitated the structural elucidation of the predominant alkylation products and provided evidence for the unusual N-7 hydrogen tautomeric form in one of them, 3-N-benzyl-6-N-acetyl-1-N-methylisoguanine (8).  相似文献   

15.
9,10-Phenanthrenedione was reacted with equimolar amounts of N,N-dimethylguanidine or creatine in 0.2 N potassium hydroxide in ethanol-water, 7:3 to obtain 2′-(dimethylamino)spiro-[9H-fluorene-9,4′-[4H]imidazol]-5′(3′H)one or N-(3′,5′-dihydro-5′-oxospiro[9H-fluorene-9,4′-[4H]imidazol]-2′-yl)-N-methylglycine, respectively. These products are the first derivatives of this ring system with 2′-amino substituents. Formation of these products accounts for the previously reported absence of fluorescence when 9,10-phenanthrenedione reacts with N,N-di-substituted guanidines.  相似文献   

16.
The syntheses of different 9-(N-phthalyl- or N-tosyl- or free aminoacyl)carbazoles and the corresponding derivatives of 3,6-dinitrocarbazoles and some derivatives of 3,6-diamino-9-(N-phthalylaminoacyl)carbazoles (II-XXXII) are described. Compounds VIII, XIII-XVII and XXIII-XXVII were found to be active against a number of microorganisms.  相似文献   

17.
Thirteen, benzo[a]phenoxazin-5-one derivatives 3a-m were synthesized from 4-nitrosoaniline hydrochlorides 1a-m and ethyl 1,3-dihydroxynaphthoate 2 and their fluorescence properties were discussed in terms of the electronic effect of substituents. A coupling reaction was carried out with 6-carbethoxy-9-N-(2-hydroxyethyl)-N-methylamino-5H-benzo[a]phenoxazin-5-one (3k) and acetyl-DL-alanine to afford N-[(6-carbethoxy-5-oxo-5H-benzo[a]phenoxazin)-9-yl]-N-methylaminoethylene acetyl-DL-alanine ester (4).  相似文献   

18.
N-(4-Nitrophenyl)oxamic acid 1 , N-(2-fluoro-4-nitrophenyl)oxamic acid 7 , N-(4-nitrophenyl)-trifluoroacetamide 3 , and N-(2-methoxy-4-nitrophenyl)trifluoroacetamide 9 are non-competitive inhibitors of Vibrio cholerae sialidase with Ki-values ranging from 2.66 to 5.18 · 10?4 M . These compounds, and the N-acetylneuraminic-acid analogues 11–13 do not inhibit the sialidase and trans-sialidase activities from Trypanosoma cruzi; nor does N-(4-nitrophenyl)oxamic acid ( 1 ) inhibit the corresponding enzyme activities from T. brucei.  相似文献   

19.
The synthesis of carbazola substituted N-acylated polyethylenimines, namely, poly[N-(9-carbazolyl)acetylethylenimine] 20 and poly[N-(2-(9-carbazolyl))propanoylethylenimine] 21 by a grafting reaction onto PEI and isomerization polymerization of the carbazole substituted 2-oxazolines is reported. A complete acylation of amino groups in PEI by the 9-carbazolylacetyl groups was achieved by the p-nitrophenyl active ester method but PEI was only partially N-acylated by the 2-(9-carbazolyl)propanoyl groups under similar reaction conditions. The carbazole substituted 2-oxazolines, namely, 2-(9-carbazolyl)methyl-2-oxazoline 18 and (R,S)-2-[1-(9-carbazolyl)]ethyl-2-oxazoline 19 , were prepared by a base induced cyclization of ß-chloroamides. The ring-opening isomerization polymerization of 18 and 19 in the molten state with a cationic initiator (dimethyl sulfate, methyl triflate, or ethylene glycol ditosylate) gave 20 and 21. Gel permeation chromatography of 20 and 21 obtained with different monomerto-initiator ratios gave evidence of a chain transfer reaction with the monomer. The polymers were characterized by elemental analyses, IR, and 1H-NMR spectroscopy.  相似文献   

20.
The synthesis of new deoxyribose nucleosides by coupling chloropurines with modified D -ribose derivatives is reported. The methyl 2-deoxy-N-methyl-3-O-(p-toluoyl)-α-D -ribofuranosiduronamide (α-D - 8 ) and the corresponding anomer β-D - 8 were synthesized starting from the commercially available 2-deoxy-D -ribose ( 1 ) (Scheme 1). Reaction of α-D - 8 with the silylated derivative of 2,6-dichloro-9H-purine ( 9 ) afforded regioselectively the N9-(2′-deoxyribonucleoside) 10 as anomeric mixture (Scheme 2), whereas β-D - 8 did not react. Glycosylation of 9 or of 6-chloro-9H-purine ( 17 ) with 1,2-di-O-acetyl-3-deoxy-N-methyl-β-D -ribofuranuronamide ( 13 ) yielded only the protected β-D -anomers 14 and 18 , respectively (Scheme 3). Subsequent deacetylation and dechlorination afforded the desired nucleosides β-D - 11 , β-D - 12,15 , and 16 . The 3′-deoxy-2-chloroadenosine derivative 15 showed the highest affinity and selectivity for adenotin binding site vs. A1 and A2A adenosine receptor subtypes.  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