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.
An efficient enantioselective synthesis of 3‐acetoxy transβ‐lactams 7a and 7b via [2+2] cycloaddition reactions of imines 4a and 4b , derived from a polycyclic aromatic amine and bicyclic chiral acid obtained from (+)‐car‐3‐ene, is described. The cycloaddition was found to be highly enantioselective, producing only trans‐(3R,4R)‐N‐azetidin‐2‐one in very good yields. This is the first report of the synthesis of enantiomerically pure transβ‐lactams 7a and 7b with a polycyclic aromatic substituent at N(1) of the azetidin ring.  相似文献   

2.
The first total synthesis of sphingolipid (2S,3R,4E)‐N2‐octadecanoyl‐4‐tetradecasphingenine ( 1a ), a natural sphingolipid isolated from Bombycis Corpus 101A, and of its styryl analogue 1b was achieved in good overall yield (Schemes 1 and 2). The key step involved the installation with (E) stereoselectivity of a long lipophilic chain or phenyl group on allyl alcohol derivative 3 via a cross‐metathesis reaction (→ 5a or 5b ). The N‐Boc protected 3 was easily accessible from (S)‐Garner aldehyde.  相似文献   

3.
It has been shown previously that the reaction of diazomethane with 5‐benzylidene‐3‐phenylrhodanine ( 1 ) in THF at ?20° occurs at the exocyclic C?C bond via cyclopropanation to give 3a and methylation to yield 4 , respectively, whereas the corresponding reaction with phenyldiazomethane in toluene at 0° leads to the cyclopropane derivative 3b exclusively. Surprisingly, under similar conditions, no reaction was observed between 1 and diphenyldiazomethane, but the 2‐diphenylmethylidene derivative 5 was formed in boiling toluene. In the present study, these results have been rationalized by calculations at the DFT B3LYP/6‐31G(d) level using PCM solvent model. In the case of diazomethane, the formation of 3a occurs via initial Michael addition, whereas 4 is formed via [3+2] cycloaddition followed by N2 elimination and H‐migration. The preferred pathway of the reaction of 1 with phenyldiazomethane is a [3+2] cycloaddition, subsequent N2 elimination and ring closure of an intermediate zwitterion to give 3b . Finally, the calculations show that the energetically most favorable reaction of 1 with diphenyldiazomethane is the initial formation of diphenylcarbene, which adds to the S‐atom to give a thiocarbonyl ylide, followed by 1,3‐dipolar electrocyclization and S‐elimination.  相似文献   

4.
A new synthesis of (Aib‐Pro)n oligopeptides (n=2, 3, and 4) via azirine coupling by using the dipeptide synthon methyl N‐(2,2‐dimethyl‐2H‐azirin‐3‐yl)‐L ‐prolinate ( 1b ; Fig. 1) is presented. The most important feature of the employed protocol is that no activation of the acid component is necessary, i.e., no additional reagents are required, and the coupling reaction is performed under mild conditions at room temperature. As an attempt to provide an answer to the question of the preferred conformation of the prepared molecules, we carried out experiments by using NMR techniques and X‐ray crystallography. For example, in the case of the hexapeptide 11 , it was possible to compare the conformations in the crystalline state and in solution. After the selective hydrolysis of the methyl ester p‐BrBz‐(Aib‐Pro)4‐OMe ( 13 ) under basic conditions, the corresponding octapeptide acid was obtained, which was then converted into the octapeptide amide p‐BrBz‐(Aib‐Pro)4‐NHC6H13 ( 15 ) by using standard coupling conditions and activating reagents (HOBt/TBTU/DIEA) of the peptide synthesis. The conformation of this compound, as well as those of the tetrapeptides 14 and 18 , was also established by X‐ray crystallography and in solution by NMR techniques. In the crystalline state, a β‐bend ribbon structure is the preferred conformation, and similar conformations are formed in solution.  相似文献   

5.
The gluco‐configured analogue 15 of nagstatin ( 1 ) and the methyl ester 14 were synthesized via condensation of the thionolactams 17 or 18 with the β‐amino ester 19 . The silyl ethers 20 and 21 resulting from 17 were desilylated to 22 and 23 ; these alcohols were directly obtained by condensing 18 and 19 . The attempted substitution of the C(8)? OH group of 22 by azide under Mitsunobu conditions led unexpectedly to the deoxygenated α‐azido esters 24 . The desired azide 25 was obtained by treating the manno‐configured alcohol 23 with diphenyl phosphorazidate. The azide was transformed to the debenzylated acetamido ester 14 that was hydrolyzed to the nagstatin analogue 15 . The imidazole‐2‐acetates 14 and 15 are nanomolar inhibitors of the N‐acetyl‐β‐glucosaminidases from Jack beans and from bovine kidney, submicromolar to micromolar inhibitors of the β‐glucosidase from Caldocellum saccharolyticum, and rather weak inhibitors of the snail β‐mannosidase. In all cases, the ester was a stronger inhibitor than the corresponding acid. As expected from their gluco‐configuration, both imidazopyridines 14 and 15 are stronger inhibitors of the βN‐acetylglucosaminidase from bovine kidney than nagstatin.  相似文献   

6.
The folate analogue, 9‐thia‐5,10‐dideazafolic acid ( 3b ), was obtained in an efficient two‐step procedure in an overall yield of 60%. The previously unknown intermediate dimethyl‐thiocarbamic acid S‐(2‐amino‐3,4‐dihydo‐4‐oxo‐pyrido[2,3‐d]pyrimidin‐6‐yl) ester ( 5 ) was prepared via the condensation of 2,6‐diamino‐3H‐pyrimidin‐4‐one and S‐(2‐malonaldehyde)‐1,1,3,3‐tetramethylthiouronium bromide ( 4 ). Compound 5 , in a one pot procedure, was deprotected using sodium hydroxide and then coupled to diethyl N‐[(4‐chloromethyl)benzoyl]‐L‐glutamate, followed by saponification of the ethyl esters to give the 9‐thia‐5,10‐dideazafolic acid ( 3b ). Compound 3b was a potent inhibitor of human 5‐aminoimidazole‐4‐carboxamide ribonucleotide transformylase (Ki of 8 ± 5 μM) and showed no inhibition of human glycinamide ribonu‐cleotide transformylase at concentrations as high as 50 μM. Compound 3b was screened by the National Cancer Institute Developmental Therapeutics Program against 60 human tumors and was found to be active against a leukemia RPMI‐8226 cell line where the LC50 was 1 μM.  相似文献   

7.
Addition of various amines to the 3,3‐bis(trifluoromethyl)acrylamides 10a and 10b gave the tripeptides 11a – 11f , mostly as mixtures of epimers (Scheme 3). The crystalline tripeptide 11f 2 was found to be the N‐terminal (2‐hydroxyethoxy)‐substituted (R,S,S)‐ester HOCH2CH2O‐D ‐Val(F6)‐MeLeu‐Ala‐OtBu by X‐ray crystallography. The C‐terminal‐protected tripeptide 11f 2 was condensed with the N‐terminus octapeptide 2b to the depsipeptide 12a which was thermally rearranged to the undecapeptide 13a (Scheme 4). The condensation of the epimeric tripeptide 11f 1 with the octapeptide 2b gave the undecapeptide 13b directly. The undecapeptides 13a and 13b were fully deprotected and cyclized to the [5‐[4,4,4,4′,4′,4′‐hexafluoro‐N‐(2‐hydroxyethoxy)‐D ‐valine]]‐ and [5‐[4,4,4,4′,4′,4′‐hexafluoro‐N‐(2‐hydroxyethoxy)‐L ‐valine]]cyclosporins 14a and 14b , respectively (Scheme 5). Rate differences observed for the thermal rearrangements of 12a to 13a and of 12b to 13b are discussed.  相似文献   

8.
β‐Amino acids are key structural elements in unnatural peptides, peptidomimetics, and many other physiologically active compounds. In view of their importance, we have developed an efficient synthetic route that provides highly enantiomerically enriched (R)‐ and (S)‐H‐β2‐htLeu‐OH via highly diastereo‐ and regioselective addition of tert‐butyl radical to enantiomerically pure N‐fumaroyloxazolidinones, followed by removal of the chiral auxiliary, Curtius rearrangement, ester hydrolysis, and catalytic hydrogenolysis.  相似文献   

9.
The synthesis of compound 2 and its derivatives 6 and 8 combining a pyrrolidine ring with an 1H‐pyrrole unit is described (Scheme 2). Their attempted usability as organocatalysts was not successful. Reacting these simple pyrrolidine derivatives with cinnamaldehyde led to the tricyclic products 3b, 9b , and 10b first (Scheme 1, Fig. 2). The final, major products were the pyrrolo‐indolizidine tricycles 3a, 9a , and 10a obtained via the iminium ion reacting intramolecularly with the nucleophilic β‐position of the 1H‐pyrrole moiety (cf. Scheme 1).  相似文献   

10.
To complete our panorama in structure–activity relationships (SARs) of sandalwood‐like alcohols derived from analogues of α‐campholenal (= (1R)‐2,2,3‐trimethylcyclopent‐3‐ene‐1‐acetaldehyde), we isomerized the epoxy‐isopropyl‐apopinene (?)‐ 2d to the corresponding unreported α‐campholenal analogue (+)‐ 4d (Scheme 1). Derived from the known 3‐demethyl‐α‐campholenal (+)‐ 4a , we prepared the saturated analogue (+)‐ 5a by hydrogenation, while the heterocyclic aldehyde (+)‐ 5b was obtained via a Bayer‐Villiger reaction from the known methyl ketone (+)‐ 6 . Oxidative hydroboration of the known α‐campholenal acetal (?)‐ 8b allowed, after subsequent oxidation of alcohol (+)‐ 9b to ketone (+)‐ 10 , and appropriate alkyl Grignard reaction, access to the 3,4‐disubstituted analogues (+)‐ 4f,g following dehydration and deprotection. (Scheme 2). Epoxidation of either (+)‐ 4b or its methyl ketone (+)‐ 4h , afforded stereoselectively the trans‐epoxy derivatives 11a,b , while the minor cis‐stereoisomer (+)‐ 12a was isolated by chromatography (trans/cis of the epoxy moiety relative to the C2 or C3 side chain). Alternatively, the corresponding trans‐epoxy alcohol or acetate 13a,b was obtained either by reduction/esterification from trans‐epoxy aldehyde (+)‐ 11a or by stereoselective epoxidation of the α‐campholenol (+)‐ 15a or of its acetate (?)‐ 15b , respectively. Their cis‐analogues were prepared starting from (+)‐ 12a . Either (+)‐ 4h or (?)‐ 11b , was submitted to a Bayer‐Villiger oxidation to afford acetate (?)‐ 16a . Since isomerizations of (?)‐ 16 lead preferentially to β‐campholene isomers, we followed a known procedure for the isomerization of (?)‐epoxyverbenone (?)‐ 2e to the norcampholenal analogue (+)‐ 19a . Reduction and subsequent protection afforded the silyl ether (?)‐ 19c , which was stereoselectively hydroborated under oxidative condition to afford the secondary alcohol (+)‐ 20c . Further oxidation and epimerization furnished the trans‐ketone (?)‐ 17a , a known intermediate of either (+)‐β‐necrodol (= (+)‐(1S,3S)‐2,2,3‐trimethyl‐4‐methylenecyclopentanemethanol; 17c ) or (+)‐(Z)‐lancifolol (= (1S,3R,4Z)‐2,2,3‐trimethyl‐4‐(4‐methylpent‐3‐enylidene)cyclopentanemethanol). Finally, hydrogenation of (+)‐ 4b gave the saturated cis‐aldehyde (+)‐ 21 , readily reduced to its corresponding alcohol (+)‐ 22a . Similarly, hydrogenation of β‐campholenol (= 2,3,3‐trimethylcyclopent‐1‐ene‐1‐ethanol) gave access via the cis‐alcohol rac‐ 23a , to the cis‐aldehyde rac‐ 24 .  相似文献   

11.
A series of new C2‐symmetric (1S,2S)‐cyclohexane‐1,2‐dicarboxamides was synthesized from (1S,2S)‐cyclohexane‐1,2‐dicarbonyl dichloride and N‐benzyl‐substituted aromatic amines, which were prepared from 2‐aminopyridine, 2‐chloroaniline, and 2‐aminophenol via imine formation with benzaldehyde and subsequent reduction with NaBH4. (1S,2S)‐N,N′‐Dibenzyl‐N,N′‐bis[2‐(benzyloxy)phenyl]cyclohexane‐1,2‐dicarboxamide was converted to (1S,2S)‐N,N′‐dibenzyl‐N,N′‐bis(2‐hydroxyphenyl)cyclohexane‐1,2‐dicarboxamide via hydrogenolysis in the presence of Pd(OH)2 on active carbon powder.  相似文献   

12.
An efficient procedure for the synthesis of N‐alkyl‐2,5‐diaryl‐1,3‐dioxol‐4‐amines 3 via a one‐pot reaction of aromatic aldehydes 2 and alkyl isocyanides 1 at room temperature in good yields is described (Scheme 1, Table).  相似文献   

13.
Methyl (22E)‐3‐oxochola‐1,4,22‐trien‐24‐oate ( 4 ; C25H34O3) is a naturally occurring steroid with unknown configuration at C(20). Starting from the (20S)‐3‐oxo‐23,24‐dinorchol‐4‐en‐22‐al ( 1a ), we prepared both diastereoisomeric methyl esters 4a and 4b by a three‐step procedure (Scheme). In the case of 4b , the initial epimerization of aldehyde 1a was followed by completion of the sequence and then separation via fractional crystallization to afford pure (20R)‐methyl ester 4a and its (20S)‐diastereomer 4b . Only the analytical data of the (20S)‐compound 4b were in good agreement with those reported for the natural product.  相似文献   

14.
Unsymmetric cyclic ketenes were generated from N‐acyl‐1,3‐thiazolidine‐2‐carboxylic acids 1a – c by means of Mukaiyama's reagent, and then reacted with imines 2a – c to the new, isomeric spiro‐β‐lactams 3 and 4 via [2+2] cycloaddition (Staudinger ketene–imine reaction; Scheme 1). The reactions were stereoselective (Table 1) and mainly afforded the spiro‐β‐lactams with a relative trans configuration. The spiro‐β‐lactams could be transformed into the corresponding monocyclic β‐lactams by means of thiazolidine ring opening or into substituted thiazolidines via hydrolysis of the β‐lactam ring.  相似文献   

15.
A simple synthesis of N2‐alkyl‐N3‐[2‐(1,3,4‐oxadiazol‐2‐yl)aryl]benzofuran‐2,3‐diamines 5 via a one‐pot four‐component reaction is described (Scheme 1). A mixture of N‐(isocyanoimino)triphenylphosphorane ( 1 ), a 2‐aminobenzoic acid 2 , a 2‐hydroxybenzaldehyde 3 , and an isocyanide 4 in absolute EtOH at room temperature undergoes a smooth reaction to afford 5 in excellent yields (Table).  相似文献   

16.
A convenient method for the synthesis of 2,4,8‐trisubstituted 1,7‐naphthyridines 6 by the reaction of (E)‐4‐(1‐aryl‐2‐methoxyethenyl)‐3‐isocyanopyridines 4 , which could be easily prepared from commercially available 3‐aminopyridine via aroylation of lithium (4‐lithiopyridin‐3‐yl)pivalamide with N‐methoxy‐N‐methylbenzamides, with excess organolithiums has been developed.  相似文献   

17.
A new series of 2,3‐disubstituted quinazolin‐4(3H)‐one derivatives was synthesized by nucleophilic attack at C(2) of the corresponding key starting material 2‐propyl‐4H‐3,1‐benzoxazin‐4‐one (Scheme 2). The reaction proceeded via amidinium salt formation (Scheme 3) rather than via an N‐acylanthranilimide. The structure of the prepared compounds were elucidated by physical and spectral data like FT‐IR, 1H‐NMR, and mass spectroscopy.  相似文献   

18.
A facile synthesis of N‐α‐Boc‐1,2‐dialkyl‐L‐histidines starting from N‐α‐trifluoroacetyl‐L‐histidine methyl ester is reported. The key steps involve direct and regiospecific N‐1(τ) ring‐alkylation of the N‐α‐trifluoroacetyl‐L‐histidine‐methyl ester by suitable alkyl iodide in the presence of NaH in DMF at ?15 °C followed by homolytic free radical C‐2 alkylation via a silver catalyzed oxidative decarboxylation of alkylcarboxylic acid in the presence of ammonium persulfate under acidic conditions. The application of newly synthesized bioimidazoles was illustrated by their incorporation into thyrotropin‐releasing hormone (TRH). The synthesized TRH analogs were evaluated in vivo for analeptic activity. We report discovery of a TRH analog, which was found to potentiate the pentobarbital‐induced sleep in vivo.  相似文献   

19.
The reaction of the 6‐substituted 1‐methyl‐4‐quinolone‐3‐carboxylates 10a , 10b with hydrazine hydrate gave the 3‐carbohydrazides 7a , 7b , respectively, whose reaction with 2‐, 3‐, and 4‐pyridinecarbaldehydes afforded the 3‐(N2‐pyridylmethylene)carbohydrazides 8a , 8b , 8c and 9a , 9b , 9c . The Curtius rearrangement of compound 7b provided the N,N′‐bis(4‐quinolon‐3‐yl)urea 14 presumably via the 3‐carboazide 11 and then 3‐isocyanate 12 . Compounds 7a , 8a , and 9a were found to possess antimalarial activity from the in vitro screening data. J. Heterocyclic Chem.,(2011).  相似文献   

20.
The diastereoselective synthesis of the N‐ and O‐protected hoprominol derivative (R,R,R)‐ 6 is described. The building up of the bicyclic O‐silylated and di(N‐tosylated) asymmetric scaffold 6 succeeded by convergent preparation of the two basic chiral azalactam units 7a and 7b and their subsequent iterative linking by a known method (Scheme 5). Both 4‐alkyl‐hexahydro‐1,5‐diazocin‐2(1H)‐ones 7a and 7b were prepared from the chiral β‐amino acid portions 10a and 10b , respectively, by application of a set of reactions (e.g., N‐alkylation of 10a , b and Sb(OEt)3‐assisted cyclization of the resulting open‐chain intermediates) already known. In comparison with the total syntheses of homaline ( 1 ) and homoprine ( 2 ), the newness of the described synthesis lies in the asymmetric approach to the difunctionalized fatty acid derivative 10b starting from (?)‐(S)‐malic acid ( 9 ) (Schemes 3 and 4). Key step in the preparation of 10b was the diastereoselective amination of the optically pure α,β‐unsaturated δ‐hydroxy homoallylic ester 14 via conjugate intramolecular aza‐Michael cyclization of the acylic δ‐(carbamoyloxy) intermediate 11 .  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