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 15 毫秒
1.
8‐Chloro‐9‐fluoro‐5‐phen­yl‐3,4,4a,5,6,10b‐hexa­hydro‐2H‐pyrano[3,2‐c]quinoline and 10‐chloro‐9‐fluoro‐5‐phen­yl‐3,4,4a,5,6,10b‐hexa­hydro‐2H‐pyrano­[3,2‐c]quinoline, both C18H17ClFNO, are diastereo­isomers, formed as the result of the imino Diels–Alder reactions of N‐benzyl­ideneanilines with 3,4‐dihydro‐2H‐pyran. The crystal structures reveal the stereochemistry of the pyran ring, which is endo/exo to the quinoline ring system formed in the cyclo­addition step. In both structures, the pyran ring adopts a chair conformation, while the nitrogen‐containing heterocyclic ring prefers a half‐chair conformation. The structures differ essentially in the relative orientation of the ring junction H atoms.  相似文献   

2.
The structures of two diastereoisomers of 9‐chloro‐8‐fluoro‐4‐phenyl‐2,3,3a,4,5,9b‐hexa­hydro­furo­[3,2‐c]­quinoline, C17H15ClFNO, are very similar. The orientation of the furan ring, as a result of its fusion to the quinoline nucleus, constitutes the significant difference between the two structures. The dihedral angles between the furan and phenyl rings are 73.4 (1) and 63.8 (1)°.  相似文献   

3.
A simple and efficient method for the cis‐selective synthesis of pyrano‐ and furano[3,2‐c]quinolines via gadolinium chloride catalyzed one‐pot aza–Diels–Alder reaction is described. Solvent conditions played a major role in affecting the diastereoselectivity of the products. © 2010 Wiley Periodicals, Inc. Heteroatom Chem 21:351–354, 2010; Published online in Wiley InterScience ( www.interscience.wiley.com ). DOI 10.1002/hc.20612  相似文献   

4.
2‐Methyl‐3H‐indoles 1 cyclize with two equivalents of ethyl malonate 2 to form 4‐hydroxy‐11H‐benzo[b]pyrano[3,2‐f]indolizin‐2,5‐diones 3, whereas 2‐mefhyl‐2,3‐dihydro‐1H‐indoles 9 give under similar conditions regioisomer 8‐hydroxy‐5‐methyl‐4,5‐dihydro‐pyrrolo[3,2,1‐ij]pyrano[3,2‐c]quinolin‐7,10‐diones 10 . The pyrone rings of 3 and 9 can be cleaved either by alkaline hydrolysis to give 7‐acetyl‐8‐hydroxy‐10H‐pyrido[1,2‐a]indol‐6‐ones 4 or 5‐acetyl‐6‐hydroxy‐2‐methyl‐1,2‐dihydro‐4H‐pyrrolo‐[3,2,1‐ij]quinolin‐4‐ones 11 , respectively. Chlorination of 3 and 9 with sulfurylchloride gives under subsequent ring opening 7‐dichloroacetyl‐8‐hydroxy‐10H‐pyrido[1,2‐a]indol‐6‐ones 5 or 5‐dichloracetyl‐6‐hydroxy‐2‐methyl‐1,2‐dihydro‐4H‐pyrrolo[3,2,1‐ij]quinolin‐4‐ones 12 . The dichloroacetyl group of 5 can be reduced with zinc to 7‐acetyl‐8‐hydroxy‐10H‐pyrido[1,2‐a]indol‐6‐ones 7. Treatment of the acetyl compounds 4, 7 and 11 with 90% sulfuric acid cleaves the acetyl group and yields 8‐hydroxy‐10H‐pyrido[1,2‐a]‐indol‐6‐ones 6 and 8 , and 6‐hydroxy‐2‐methyl‐1,2‐dihydro‐4H‐pyrrolo[3,2,1‐ij]quinolin‐4‐ones 13 . Reaction of dichloroacetyl compounds 12 with sodium azide yields 6‐hydroxy‐2‐methyl‐5‐(1H‐tetrazol‐5‐ylcarbonyl)‐1,2‐dihydro‐4H‐pyrrolo[3,2,1‐ij]quinolin‐4‐ones 14 via intermediate geminal diazides.  相似文献   

5.
A mild and efficient method for the synthesis of cyclopenta[c]naphtho[2,3‐f]quinoline and cyclopenta[c]pyrazolo[4,3‐f]quinoline derivatives via an imino Diels–Alder reaction of aromatic aldehyde, anthracen‐2‐amine, or 1H‐indazol‐5‐amine and cyclopentanone catalyzed by iodine is described. This novel procedure has the advantages of mild reaction condition, high yields, and metal‐free catalyst.  相似文献   

6.
A synthesis of pyrrolo[3,2‐c]quinolines substituted in the 7‐ and 8‐ positions by methoxy groups and in the 3‐ position by an amido group is described. The structures were designed to have a crescent shape, a planar fused cyclic moiety with two ortho methoxy groups and ionisable amino or amidinic group at pH 7.  相似文献   

7.
Treatment of 2‐hydroxy‐, 2‐mercapto‐, and 2‐ethoxycarbonylamino‐benzonitriles 12 with 2‐fluoro‐ or 2‐nitrophenacylbromides 13 under alkaline conditions provided the corresponding benzofuran, benzothiophene, and indole intermediates 10 , respectivelly. Nucleophilic cyclization of these compounds led to the corresponding tetracyclic quinolinones 7a, 7b , and 3. Denitrocyclization reaction of compounds 10 (R = NO2) was found especially useful. Compounds 7a, 7b , and 3 were converted to their chloro derivatives 14a‐c , which were reduced with hydrogen and a catalyst to the corresponding compounds 8a, 8b , and 2. The presented pathway represents a new method of preparation of quindoline 2 and its O and S analogs 8. Chloro derivatives 14 are reactive enough to provide the corresponding methoxy derivatives 15 and dimethylamino derivatives 16. Methylation of compounds 7a and 7b with iodomethane providing mixtures of major N‐methyl derivatives 17 and minor O‐methyl derivatives 15 were also studied.  相似文献   

8.
9.
The density functional theory level (B3LYP/6‐311G**) computations of the Diels–Alder (DA) reactions of 5,6‐dihydrothiazolo[3,2‐d][1,4,2]‐diazaphospholes with 1,3‐butadiene and with isoprene confirm pericyclic mechanism via asynchronous transition states. The aromatic character of the transition states is established by negative nucleus independent chemical shift (NICS) values falling in the range from −14 to −16. Integration of the dienophilic >CP functionality in the 6π aromatic azaphosphole ring raises the activation energy barrier (B3LYP/6‐311++G**//B3LYP/6‐311G**) compared to that for the DA reaction of the acyclic phosphaethene, but it is lower than the activation energy barrier for the DA reaction of the corresponding 10π aromatic system, thiazolo[3,2‐d][1,4,2]diazaphospholes. The experimentally observed stereo‐ and regioselectivities in the reactions can be accounted on the basis of secondary molecular orbital (SMO) interactions detected in the respective transition structures. The attachment of an electron‐withdrawing group to the dienophilic moiety enhances both stereo‐ and regio‐ selectivities which agree well with the experimental values. Solvent (toluene) effect studied with polarizable continuum model (PCM) indicates that the stereo‐ and regioselectivities are not affected by the solvent. © 2006 Wiley Periodicals, Inc. Heteroatom Chem 17:402–410, 2006; Published online in Wiley InterScience ( www.interscience.wiley.com ). DOI 10.1002/hc.20196  相似文献   

10.
The imino Diels–Alder reaction is an efficient method for the synthesis of aza‐heterocycles. While different stereo‐ and enantioselective inverse‐electron‐demand imino Diels–Alder (IEDIDA) reactions have been reported before, IEDIDA reactions including electron‐deficient dienes are unprecedented. The first enantioselective IEDIDA reaction between electron‐poor chromone dienes and cyclic imines, catalyzed by zinc/binol complexes is described. The novel reaction provides a facile entry to a natural product inspired collection of ring‐fused quinolizines including a potent modulator of mitosis.  相似文献   

11.
1 mol% of catalyst is sufficient : The hetero Diels–Alder reaction of α-imino esters 1 with activated conjugated dienes 2 (R=H, Me) needs only 1 mol% of a 2,2′-bis(diarylphosphanyl)-1,1′-binaphthyl (BINAP) copper(I ) complex as the catalyst to generate the adducts 3 in good yields and with enantioselectivities up to 96%. The reaction can also be carried out on gram scale! Tos=H3CC6H4SO2; TMS=Me3Si.  相似文献   

12.
13.
Three‐component reaction of 4‐hydroxy‐2H‐benzo[h]chromen‐2‐one, aromatic aldehydes, and malononitrile in the presence of 1,4‐diazabicyclo[2.2.2]octane (DABCO) in ethanol at room temperature affords good yields of novel dihyrobenzo[h]pyrano[3,2‐c]chromene derivatives. The synthesized compounds examined by MTT assays for cytotoxic activity in two human cancer cell lines (MOLT‐4, HL‐60). Most of the evaluated compounds showed low inhibitory activity against tumor cell line at micromolar concentrations.  相似文献   

14.
1‐Ethyl‐4‐hydroxy‐2‐oxo‐1,2‐dihydroquinoline‐3‐carbaldehyde ( 1 ) was annulated using malonic acid and/or its ethyl ester to furnish pyrano[3,2‐c]quinoline‐3‐carboxylic acid 2 and its ester 3 . Interconversions between acid 2 and ester 3 were successfully carried out. The anticipated pyrano[3,2‐c]quinoline‐3‐carboxamides 5–12 were conveniently attained via condensation of ester 3 with the proper amine. Surprisingly, treatment of ester 3 with dimethylformamide (DMF) in acidic media led to the carboxamide 5 . All attempts to convert ester 3 to its corresponding acid hydrazides by interaction with the proper hydrazine derivative led to formation of pyrazolidinediones 15 and 17 . Ester 3 underwent cyclo‐condensation with malononitrile dimer affording pyrido[3′,4′:5,6]pyrano[3,2‐c]quinoline derivative 18 . The new compounds revealed significant antioxidant effect, which suggests that most of them are possible potent antioxidant agents.  相似文献   

15.
Diels–Alder reactions employing 1,2‐azaborine heterocycles as 1,3‐dienes are reported. Carbocyclic compounds with high stereochemical and functional complexity are produced, as exemplified by the straightforward two‐step synthesis of an amino allyl boronic ester bearing four contiguous stereocenters as a single diastereomer. Whereas electron‐deficient dienophiles undergo irreversible Diels–Alder reactions, a reversible Diels–Alder reaction with the less electron‐deficient methyl acrylate is observed. Both the N and the B substituent of the 1,2‐azaborine exert significant influence on the [4+2] cycloaddition reactivity as well as the aromatic character of the heterocycle. The experimentally determined thermodynamic parameters of the reversible Diels–Alder reaction between 1,2‐azaborines and methyl acrylate correlate with aromaticity trends and place 1,2‐azaborines approximately between furan and thiophene on the aromaticity scale.  相似文献   

16.
Synthesis of novel 3‐(ethoxymethylene)‐pyrano[3,2‐c]quinolinone and pyrano[3,2‐c]quinoline‐3‐carboxaldehyde was accomplished efficiently via a simple method. These two scaffolds were used as precursors to afford new biologically interesting products in good yield and short reaction times. The chemical reactivity of ethoxy methylene 2 and carboxaldehyde 3 toward different nucleophilic reagents was studied. Structures of the new synthesized compounds were elucidated by their analytical and spectral data.  相似文献   

17.
18.
Among concerted cycloadditions, the Diels–Alder reaction is the grand old classic, which is usually achieved with acid catalysis. In this report, hydroxypyrones, oxa‐, and thiazolones are explored because they provide access to anionic dienes. Their [4+2] cycloaddition with cyclic and acyclic dienophiles, such as maleimides and fumarates, affords bicyclic products with four new stereogenic centers. Bifunctional anion–π catalysts composed of amine bases next to the π surface of naphthalenediimides (NDIs) are shown to selectively stabilize the “open”, fully accessible anionic exo transition state on the π‐acidic aromatic surface. Our results also include reactivities that are hard to access with conventional organocatalysts, such as the exo ‐specific and highly enantioselective Diels–Alder reaction of thiazolones and maleimides with complete suppression of the otherwise dominant Michael addition. With increasing π acidity of the anion–π catalysts, the rates, chemo‐, diastereo‐, and enantioselectivities increase consistently.  相似文献   

19.
20.
Synthetic methods have been developed to prepare pyrano[2,3‐c]pyrazoles with various substituents at ring positions 1, 3, and 6. The 1H‐ and 13C‐NMR properties of these products and their precursors are presented and discussed. J. Heterocyclic Chem., (2011).  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