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 46 毫秒
1.
Reaction of 2-formyl-2-(2,3-O-isopropylidene-5-O-trityl-D-ribofuranosyl)acetonitrile (VII) with semicarbazide hydrochloride followed by sodium ethoxide treatment afforded an α,β-mixture of 3-amino-2N-carbamoyl-4-(2,3-O-isopropylidene-5-O-trityl-D-ribofuranosyl)pyrazole (IX). Conversion of IX to 4-oxo-8-(2,3-O-isopropylidene-5-O-trityl-D-ribofuranosyl)-3H-pyrazolo[1,5-a]-1,3,5-triazine (XIII) was achieved by treatment of IX with ethylorthoformate. The β-isomer IXb gave only the β-isomer XIIIb, and the α-isomer IXa was converted exclusively into the α-isomer XIIIa. Upon deprotection with 3% n-butanolic hydrogen chloride, both IXa and IXb gave the same mixture of the α- and β-isomers of 3-amino-2N-carbamoyl-4-(D-ribosyl)pyrazole, which were separated by chromatography. The syntheses of the hitherto unknown compounds, 3-amino-2N-carbamoylpyrazole (IVa) and its 4-methyl analog (IVb) are also reported. Experimental details of the synthesis of 3-amino-4-(2,3-O-isopropylidene-5-O-trityl-β-D-ribofuranosyl)pyrazole (XIIb), an important intermediate for “purine-like” C-nucleosides, are also described.  相似文献   

2.
α-(Aminomethylene)-9-(methoxymethyl)-9H-purine-6-acetamide and the ethyl acetate, 3 and 8 , have been synthesized by catalytic hydrogenation of 6-cyanomethylene-9-methoxymethylpurine derivatives 2 and 7 which were obtained by the substitution of 6-chloro-9-(methoxymethyl)purine ( 1 ) with α-cyanoacetamide and ethyl cyanoacetate, respectively. Substitution of 3 and 8 with amines gave the corresponding N-substituted α-(aminomethylene)-9-(methoxymethyl)-9H-purine-6-acetamide and the ethyl acetate 4 and 10 . Reaction of 3 with piperidine gave 9-(methoxymethyl)-9H-purine-6-acetamide ( 5 ).  相似文献   

3.
Preparation of the Enantiomerically Pure cis- and trans-Configurated 2-(tert-Butyl)-3-methylimidazolidin-4-ones from the Amino Acids (S)-Alanine, (S)-Phenylalanine, (R)-Phenylglycine, (S)-Methionine, and (S)-Valine In contrast to α-hydroxy and α-mercapto carboxylic acids, simple α-amino acids do not form acetal-type derivatives ( 2 , X = NH) with pivalaldehyde. For the generation of amino-acid-derived chiral, nonracemic enolates (cf. 3 ), and hence, for the α-alkylation of amino acids without racemization and without an external chiral auxiliary, the imidazolidinones 12–14 were prepared diastereoselectively. To this end, the methyl or ethyl esters of amino-acid hydrochlorides were first converted to N-methylamides of amino acids which in turn were condensed with pivalaldehyde to give (neopentylidenamino)amides ( 11 ). These Schiff bases could be cyclized either to trans-or to cis-imidazolidinones ( 12, 14 and 13 , respectively), which were obtained in enantiomerically pure form after recrystallization. The enantiomeric purities were confirmed by HPLC with chiral stationary phases or by 1H-NMR spectroscopy in the presence of chiral shift reagents. The configurations (cis, trans) were assigned by NOE measurements on 300- or 360-MHz 1H-NMR spectrometers.  相似文献   

4.
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   

5.
3-(Dimethylamino)-2,2-dimethyl-2H-azirine as an Aib Equivalent; Synthesis of Aib Oligopeptides 3-(Dimethylamino)-2,2-dimethyl-2H-azirine ( 1 ) reacts with carboxylic acids at 0–25° to give 2-acylamino-N,N,2-trimethylpropionamides ( = 2-acylamino-N,N-dimethylisobutyramide, acyl-Aib-NMe2) in excellent yields (Scheme 2 and 3). Examples of α-amino-, α-hydroxy-, and α-mercapto-carboxylic acids are given. On treatment with HCl in toluene, the terminal dimethylamide group is selectively converted to the corresponding carboxylic acid (→acyl-Aib) via an amide cleavage (Scheme 4 and 5); 1,3-oxazol-5(4H)-ones are intermediates of this amide hydrolysis. This reaction sequence has been used for the extension of peptide chains (Scheme 6). The synthesis of Aib-oligopeptides using this methodology is described (Scheme 8).  相似文献   

6.
3-(Dimethylamino)-2,2-dimethyl-2H,-azirine as an α-Aminoisobutyric-Acid (Aib) Equivalent: Cyclic Depsipeptides via Direct Amid Cyclization In MeCN at room temperature, 3-(dimethylamino)-2,2-dimethyl-2H-azirine ( 1 ) and α-hydroxycarboxylic acids react to give diamides of type 8 (Scheme 3). Selective cleavage of the terminal N,N-dimethylcarboxamide group in MeCN/H2O leads to the corresponding carboxylic acids 13 (Scheme 4). In toluene/Ph SH , phenyl thioesters of type 11 are formed (see also Scheme 5). Starting with diamides 8 , the formation of morpholin-2,5- diones 10 has been achieved either by direct amide cyclization via intermediate 1,3-oxazol-5(4H)-ones 9 or via base-catalyzed cyclization of the phenyl thioesters 11 (Scheme 3). Reaction of carboxylic acids with 1 , followed by selective amide hydrolysis, has been used for the construction of peptides from α-hydroxy carboxylic acids and repetitive α-aminoisobutyric-acid (Aib) units (Scheme 4). Cyclization of 14a, 17a , and 20a with HCI in toluene at 100° gave the 9-, 12-, and 15-membered cyclic depsipeptides 15, 18 , and 21 , respectively.  相似文献   

7.
(E)-1-Benzotriazolyl-3-(phenylsulphonyl)-1-propene ( 6 ) has been synthesized and its alkylation was studied. The results showed that the phenylsulphonyl group is a more powerful α-directing group than the benzotriazolyl group in the corresponding 1,3-dihetero-stabilized allyl anion.  相似文献   

8.
4-(4-PhenyI-3-pyrazolyl)-4H-1,2,4-triazoles and 4-phenyl-5-(4H-1,2,4-triazol4-yl)-3-pyrazolols were prepared by the reaction of formylhydrazine on α-phenyl-α-cyanoacetaldehydes and ethyl α-phenyl-α-cyanoacetates.  相似文献   

9.
Enantiomerically pure cis- and trans-5-alkyl-1-benzoyl-2-(tert-butyl)-3-methylimidazolidin-4-ones ( 1, 2, 11, 15, 16 ) and trans-2-(tert-butyl)-3-methyl-5-phenylimidazolidin-4-one ( 20 ), readily available from (S)-alanine, (S)-valine, (S)-methionine, and (R)-phenylglycine are deprotonated to chiral enolates (cf. 3, 4, 12, 21 ). Diastereoselective alkylation of these enolates to 5,5-dialkyl- or 5-alkyl-5-arylimidazolidinones ( 5, 6, 9, 10, 13a-d, 17, 18, 22 ) and hydrolysis give α-alkyl-α-amino acids such as (R)- and (S)-α-methyldopa ( 7 and 8a , resp.), (S)-α-methylvaline ( 14 ), and (R)-α-methyl-methionine ( 19 ). The configuration of the products is proved by chemical correlation and by NOE 1H-NMR measurements (see 23, 24 ). In the overall process, a simple, enantiomerically pure α-amino acid can be α-alkylated with retention or with inversion of configuration through pivaladehyde acetal derivatives. Since no chiral auxiliary is required, the process is coined ‘self-reproduction of a center of chirality’. The method is compared with other α-alkylations of amino acids occurring without racemization. The importance of enantiomerically pure, α-branched α-amino acids as synthetic intermediates and for the preparation of biologically active compounds is discussed.  相似文献   

10.
The methanesulfonates of (α-(4-chlorophenyl)-α-[1-(2-chlorophenyl)ethenyl]-1H-1,2,4-triazole-1-ethanol and α-[1-(2-chlorophenyl)ethenyl]-α-(2,4-difluorophenyl)-1H-1,2,4-triazole-1-ethanol ( 1a, b ) are orally effective α-styryl carbinol derivatives developed for the treatment and prevention of systemic fungal infections. Practical new processes amenable for the large-scale production of these compounds are described. Of note is the selection of dichlorostyrene as a convenient precursor of the styryl portion, modification of a sensitive Grignard addition into a realistic preparative reaction and the use of 1,2,4-triazole simultaneously as a base transfer agent and nucleophile.  相似文献   

11.
Active targeting strategies are currently being extensively investigated in order to enhance the selectivity of photodynamic therapy. The aim of the present research was to evaluate whether the external decoration of nanopolymeric carriers with targeting peptides could add more value to a photosensitizer formulation and increase antitumor therapeutic efficacy and selectivity. To this end, we assessed PLGA-PLA-PEG nanoparticles (NPs) covalently attached to a hydrophilic photosensitizer 5-[4-azidophenyl]-10,15,20-tri-(N-methyl-4-pyridinium)porphyrinato zinc (II) trichloride (ZnTriMPyP) and also to c(RGDfK) peptides, in order to target αvβ3 integrin-expressing cells. In vitro phototoxicity investigations showed that the ZnTriMPyP-PLGA-PLA-PEG-c(RGDfK) nanosystem is effective at submicromolar concentrations, is devoid of dark toxicity, successfully targets αvβ3 integrin-expressing cells and is 10-fold more potent than related nanosystems where the PS is occluded instead of covalently bound.  相似文献   

12.
(R,R)-Butanediol (dichloromethyl)boronate ( 1 ) with 1 equiv. allylmagnesium halide yields (R,R)-2,3-butanediol (1S)-(1-chloro-3-butenyl)boronate ( 3 ) together with the diallylated product (R,R)-2,3-butanediol (1-allyl-3-butenyl)boronate ( 4 ). The formation of 4 is unprecedented in reactions of α-chloroboronic esters with Grignard reagents. With methylmagnesium bromide 3 yielded (R,R)-2,3-butanediol (1S)-(1-methyl-3-butenyl)boronate ( 5 ), which failed to hydrolyze with water. Hydrolysis of 3 yielded impure α-chloroboronic acid, which was esterified with pinacol and treated with methylmagnesium bromide to form 6 , which with (dichloromethyl)lithium followed by methylmagnesium bromide yielded diastereomeric boronic esters 7 and 8 . Oxidation by hydrogen peroxide yielded (2S,3S)- and (2R,3S)-3-methyl-5-hexen-2-ol ( 9 and 10 , ees unknown). Treatment of (s)-pinanediol allylboronate ( 11 ) with (dichloromethyl)lithium at −100°C followed by zinc chloride at up to 25°C has proceeded in the normal way to form (s)-pinanediol (1S)-(1-chloro-3-butenyl)-boronate ( 12 ), which has been elaborated via 13 , 14 , and 15 to (2S,3S)-3-methyl-5-hexen-2-ol ( 9 ) in 95% de.  相似文献   

13.
Nine pairs of isomeric 2,4,6-tris(halophenyl)-1,3,5-trithianes have been assayed in the crude state by n.m.r. techniques, and separated and purified by chromatography. Contrary to previous reports, the α-(cis, trans)-isomers are the major products in most cases. These compounds are shown to exist as puckered chair trithiane structures, even in the more hindered α-(cis, trans)-o-halophenyl cases, by the clear resolution of axial and equatorial trithiane ring protons in a ratio of 2:1. An o-halogen on an axial phenyl group in the α-isomers causes the aromatic group to exert an anisotropic deshielding effect on adjacent axial protons, so as to cause the axial and equatorial proton peaks to appear as a singlet in some solvents. Melting point differences, in several cases quite large, from those previously reported have been observed for six of the eighteen triaryltrithianes reported.  相似文献   

14.
An efficient synthesis of the potent and orally active 5-HT1A agonists, (R)-(+)- and (S)-(-)-1-formyl-6,7,8,9-tetrahydro-N,N-dipropyl-3H-benz[e]indol-8-amines 1a and 1b , is described. This synthesis was accomplished in twelve steps from commercially available 1,5,6,7-tetrahydro-4H-indol-4-one ( 5 ). The key step involved a regio-controlled Friedel-Crafts acylation of 1-(p-toluenesulfonyl)indol-4-acetyl chloride with ethylene to yield a versatile synthon, 3-(p-toluenesulfonyl)-6,7,8,9-tetrahydro-3H-benz[e]indol-8-one ( 10 ). Subsequent coupling of this ketone with chiral α-methylbenzylamine under reductive amination conditions yielded a mixture of diastereomers. These diastereomers were efficiently separated by either chromatography or fractional recrystallization of the derived hydrochloride salts. Debenzylation of the pure diastereomers was followed by alkylation and formylation to yield (R)-(+)- and (S)-(-)-enantiomers 1a and 1b with >99% purity.  相似文献   

15.
An efficient synthesis of enantiomerically pure (R)- and (S)-2-(aminomethyl)alanine ((R)- and (S)-Ama) 1a and (R)- and (S)-2-(aminomethyl)leucine ((R)- and (S)-Aml) 1b is described (Schemes 1 and 2). Resolution of the racemic amino acids was achieved using L -phenylalanine cyclohexylamide ( 2 ) as chiral auxiliary. The free amino acids 1a, b were converted to the Nα-Boc,Nγ-Z-protected derivatives 11a, b (Scheme 3) ready for incorporation into peptides. Based on the three crystal structures of the diastereoisomeric peptides 8a, 8b , and 9b , the absolute configurations in both series were determined. β-Turn type-I geometries were observed for structures 8b and 9b , whereas 8a crystallized in an extended backbone conformation.  相似文献   

16.
(R)- and (S)-α-ionone ((R)- and (S)- 1 , resp.) were prepared from (R)- and (S)-α-damascone ((R)- and (S)- 3 , resp.) without racemization in 48% yield employing a new enone transposition. The described transposition is complementary to existing methods whose application is often prohibited by the structural requirements of the substrate. The now easily accessible α-ionones of desired absolute configuration are useful as chiral building blocks for terpenoid synthesis.  相似文献   

17.
Cathinones belong to a group of compounds of great interest in the new psychoactive substances (NPS) market. Constant changes to the chemical structure made by the producers of these compounds require a quick reaction from analytical laboratories in ascertaining their characteristics. In this article, three cathinone derivatives were characterized by X-ray crystallography. The investigated compounds were confirmed as: 1-[1-(4-methylphenyl)-1-oxohexan-2-yl]pyrrolidin-1-ium chloride ( 1 , C17H26NO+·Cl?, the hydrochloride of 4-MPHP), 1-(4-methyl-1-oxo-1-phenylpentan-2-yl)pyrrolidin-1-ium chloride ( 2 ; C16H24NO+·Cl?, the hydrochloride of α-PiHP) and methyl[1-(4-methylphenyl)-1-oxopentan-2-yl]azanium chloride ( 3 ; C13H20NO+·Cl?, the hydrochloride of 4-MPD). All the salts crystallize in a monoclinic space group: 1 and 2 in P21/c, and 3 in P21/n. To the best of our knowledge, this study provides the first detailed and comprehensive crystallographic data on salts 1 – 3 .  相似文献   

18.
5-(1-Adamantyloxy)-2H-pyrrole-2 one has been homopolymerized and copolymerized with a variety of comonomers. Polymerization was conducted in chloroform solutions with α,α′-azobisisobutyronitrile initiator. Evidence of polymerization was achieved through infrared and NMR spectra and elemental analysis. Moderate molecular weights were achieved as determined by inherent viscosity measurements and gel-permeation chromatography. Photolysis of the polymers with ultraviolet radiation induces a photochemical rearrangement resulting in the formation of isocyanate functions. A proposed mechanism suggests α-cleavage of the carbonyl to give a 1,5-diradical which rearranges to a 1,3-diradical with subsequent ring closure to give a polymer with cyclopropyl isocyanate moieties in the backbone. DSC data show all polymers to display intense exothermic activity at temperatures near 200°C on initial heating and glass transition temperatures between 194 and 245.5°C on subsequent heating. Thermolysis of the homopolymer causes rearrangement to poly[N-(1-adamantyl) maleimide]. Reactivity ratios were determined for the systems styrene (M1) and 5-(1-adamantyloxy)-2H-pyrrole-2-one (M2) (r1 = 0.06, r2 = 0.07) and methyl methacrylate (M1) and 5-(1-adamantyloxy)-2H-pyrrole-2-one(M2) (r1 = 0.35, r2 = 0.70). Q and e values for 5-(1-adamantyloxy)-2H-pyrrole-2-one are 3.40 and 1.59, respectively.  相似文献   

19.
N-substituted 4-(2-morpholinyl)indoles were prepared from 1-(t-butoxycarbonyl)-4-acetylindole (7) which was itself prepared from 4-cyanoindole. Bromination of ketone 7, followed by reaction with amines and subsequent sodium borohydride reduction, gave amino alcohols. These were converted to α-chloro amides that were cyclized to lactams. Lithium aluminum hydride reduction served both to remove the t-BOC protecting group and to reduce the lactams to the 4-(2-morpholinyl) indoles.  相似文献   

20.
α-(Alkoxymethyl) acrylates, such as methyl α-(phenoxymethyl) acrylate, benzyl α-(methoxymethyl)acrylate (BMMA), benzyl α-(benzyloxymethyl)acrylate, and benzyl α-(tert-butoxymethyl)acrylate, were synthesized, and their polymerizability and the stereoregularity of the polymers obtained by radical and anionic methods were investigated. The radically obtained polymers were found to be atactic by 13C- and 1H-NMR analyses, but the polymers obtained with lithium reagents in toluene at −78°C were highly isotactic. Further, it is noteworthy that isotactic polymers were also produced with lithium reagents even in tetrahydrofuran. Effects of polymerization temperature and counter cation on stereoregularity were clearly observed in the polymerization of BMMA, and a potassium reagent afforded an almost atactic polymer. © 1997 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 35: 721–726, 1997  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