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 109 毫秒
1.
Starting from 1‐[(tert‐butoxy)carbonyl]piperidine‐4‐carboxylic acid and 2‐bromoaniline, the spiro[indole‐3,4′‐piperidin]‐2‐one system was obtained in three high‐yielding steps: anilide formation, N(1)‐protection, and intramolecular cyclization under Pd catalysis as the key reaction. The preparation of the corresponding 2‐bromoanilide was studied. In extension, the same sequence was developed with 4‐methyl‐ and 4‐nitro‐2‐bromoaniline. In the key step, the NO2 group led to a rather diminished yield. The transformation of the protected spiro[indole‐3,4′‐piperidin]‐2‐one to the corresponding unprotected dihydroindoles is discussed.  相似文献   

2.
The structural information gained from the study of the chiral building block (R)‐(?)‐4‐(3,4‐di­chloro­phenyl)‐4‐(2‐pyridyl)­butanoic acid–l ‐(?)‐ephedrine [methyl(1‐hydroxy‐1‐phenyl­prop‐2‐yl)ammon­ium 4‐(3,4‐di­chloro­phenyl)‐4‐(2‐pyrid­yl)but­an­oate], C10H16NO+·C15H12Cl2NO2?, can be used to deduce the absolute configuration of highly potent arpromidine‐type hist­amine H2 receptor agonists, as the chiral butanoic acid can be converted to (R)‐(?)‐3‐(3,4‐di­chloro­phenyl)‐3‐(2‐pyridyl)­propyl­amine and to the corresponding R‐configured arpromidine analogue.  相似文献   

3.
A facile one‐pot, three‐component protocol for the synthesis of novel spiro[3H‐indole‐3,2′‐thiazolidine]‐2,4′(1H)‐diones by condensing 1H‐indole‐2,3‐diones, 4H‐1,2,4‐triazol‐4‐amine and 2‐sulfanylpropanoic acid in [bmim]PF6 (1‐butyl‐3‐methyl‐1H‐imidazolium hexafluorophosphate) as a recyclable ionic‐liquid solvent gave good to excellent yields in the absence of any catalyst (Scheme 1 and Table 2). The advantages of this protocol over conventional methods are the mild reaction conditions, the high product yields, a shorter reaction time, as well as the eco‐friendly conditions.  相似文献   

4.
Unusual course of the reaction was revealed on the oxidation of functionally substituted acridine containing the activated methyl groups by well‐known oxidants, such as selenous acid and selenium(IV) oxide. Treatment of 2‐methoxy‐4,9‐dimethyl‐1‐nitroacridine ( 5 ) with an excess of H2SeO3 in boiling ethanol gave a mixture of the normal reaction products, 2‐methoxy‐4‐methyl‐1‐nitro‐9‐formylacridine ( 11 ) (isolated yield 29%) and 2‐methoxy‐4‐methyl‐1‐nitroacridine‐9‐carboxylic acid ( 12 ) (36%), together with an abnormal product, 3‐methoxy‐5‐methyl‐1H‐selenopheno[2,3,4‐k,l]acridine‐1‐one ( 13 ) (21%), which is the first example of a new seleno‐containing ring system. With the use of SeO2 the yield of selenolactone 13 was much lower. J. Heterocyclic Chem., 2011.  相似文献   

5.
Three new compounds, CuL, CuL′, and Cu2O2L′′2 (H2L=3′‐[(E)‐{[(1S,2S)‐2‐aminocyclohexyl]imino}methyl]‐4′‐hydroxy‐4‐biphenylcarboxlic acid, H2L′=3′‐[(E)‐{[(1S,2S)‐2‐aminocyclohexyl]imino}methyl]‐4′‐hydroxy‐5′‐nitro‐4‐biphenylcarboxlic acid, H2L′′=3′‐(N,N‐dimethylamino methyl)‐4′‐hydroxy‐4‐biphenylcarboxlic acid), were selectively synthesized through a controlled in situ ligand reaction system mediated by copper(II) nitrate and H2L. Selective nitration was achieved by using different solvent mixtures under relatively mild conditions, and an interesting and economical reductive amination system in DMF/EtOH/H2O was also found. All crystal structures were determined by single‐crystal X‐ray diffraction analysis. Both CuL and CuL′ display chiral 1D chain structures, whereas Cu2O2L′′2 possesses a structure with 13×16 Å channels and a free volume of 41.4 %. The possible mechanisms involved in this in situ ligand‐controlled reaction system are discussed in detail.  相似文献   

6.
The structures of new oxaindane spiropyrans derived from 7‐hydroxy‐3′,3′‐dimethyl‐3′H‐spiro[chromene‐2,1′‐isobenzofuran]‐8‐carbaldehyde (SP1), namely N‐benzyl‐2‐[(7‐hydroxy‐3′,3′‐dimethyl‐3′H‐spiro[chromene‐2,1′‐isobenzofuran]‐8‐yl)methylidene]hydrazinecarbothioamide, C27H25N3O3S, (I), at 120 (2) K, and N′‐[(7‐hydroxy‐3′,3′‐dimethyl‐3′H‐spiro[chromene‐2,1′‐isobenzofuran]‐8‐yl)methylidene]‐4‐methylbenzohydrazide acetone monosolvate, C27H24N2O4·C3H6O, (II), at 100 (2) K, are reported. The photochromically active Cspiro—O bond length in (I) is close to that in the parent compound (SP1), and in (II) it is shorter. In (I), centrosymmetric pairs of molecules are bound by two equivalent N—H...S hydrogen bonds, forming an eight‐membered ring with two donors and two acceptors.  相似文献   

7.
The reaction between homophthalic anhydride and cyclohexanone was examined both in the presence of DMAP or BF3·Et2O complex as a catalyst. The latter yielded (±)‐1‐oxo‐1H‐spiro[benzo[c]pyran‐3(4H), 1′‐cyclohexane]‐4‐carboxylic acid ( 3 ) in a higher yield (82 %). A series of new (±)‐4‐(N,N‐disubstituted‐1‐carbamoyl)‐1H‐spiro[benzo[c]pyran‐3(4H),1′‐cyclohexane]‐1‐ones ( 5a‐h ) were synthesized from the parent acid 3 by a two‐step reaction. Differentiating microbial screening was performed for most of the synthesized compounds against twelve microorganisms belonging to different taxonomic groups. The spiro acid 3 was active against all bacterial strains with MIC ≥ 20 μg/ml against B. subtillis and P. vulgaris. E. coli was the most sensitive strain to the antibacterial effect of the tested compounds.  相似文献   

8.
In the presence of a catalytic amount of an imidodiphosphoric acid, enantioselective desymmetrization of bicyclic bislactones by reaction with alcohols took place smoothly to afford enantiomerically enriched monoacids having an all‐carbon stereogenic center. Concise catalytic enantioselective syntheses of both (?)‐rhazinilam and (?)‐leucomidine B were subsequently developed using (S)‐methyl 4‐ethyl‐4‐formylpimelate monoacid as a common starting material.  相似文献   

9.
The crystal structure of methyl α‐d ‐mannopyranosyl‐(1→3)‐2‐O‐acetyl‐β‐d ‐mannopyranoside monohydrate, C15H26O12·H2O, ( II ), has been determined and the structural parameters for its constituent α‐d ‐mannopyranosyl residue compared with those for methyl α‐d ‐mannopyranoside. Mono‐O‐acetylation appears to promote the crystallization of ( II ), inferred from the difficulty in crystallizing methyl α‐d ‐mannopyranosyl‐(1→3)‐β‐d ‐mannopyranoside despite repeated attempts. The conformational properties of the O‐acetyl side chain in ( II ) are similar to those observed in recent studies of peracetylated mannose‐containing oligosaccharides, having a preferred geometry in which the C2—H2 bond eclipses the C=O bond of the acetyl group. The C2—O2 bond in ( II ) elongates by ~0.02 Å upon O‐acetylation. The phi (?) and psi (ψ) torsion angles that dictate the conformation of the internal O‐glycosidic linkage in ( II ) are similar to those determined recently in aqueous solution by NMR spectroscopy for unacetylated ( II ) using the statistical program MA′AT, with a greater disparity found for ψ (Δ = ~16°) than for ? (Δ = ~6°).  相似文献   

10.
(?)‐Hybridalactone ( 1 ) is a marine eicosanoid isolated from the red alga Laurencia hybrida. This natural product contains cyclopropane, cyclopentane, 13‐membered macrolactone and epoxide ring systems incorporating seven stereogenic centers. Moreover, this compound has an acid‐labile skipped Z,Z‐diene motif. In this paper, we report on the total synthesis of (?)‐hybridalactone ( 1 ). The unique eicosanoid (?)‐hybridalactone ( 1 ) was synthesized starting from optically active γ‐butyrolactone 2 in a linear sequence comprising 21 steps with an overall yield of 21.9 %. A key step in the synthesis of (?)‐hybridalactone ( 1 ) is the methyl phenylsulfonylacetate‐mediated one‐pot synthesis of the cis‐cyclopropane‐γ‐lactone derivative. This reaction provided an efficient and stereoselective access to cis‐cyclopropane‐γ‐lactone 12 . Further elaboration of the latter compounds through desulfonylation, epoxidation, oxidation, Wittig olefination and Shiina macrolactonization afforded (?)‐hybridalactone.  相似文献   

11.
The reaction of bi­phenyl‐2,2′‐diol with 1,1,2,2‐tetrakis­(bromo­methyl)­cyclo­propane leads to two products, namely a propellane‐type compound and a di­spiro‐type compound. The molecular structure of 4,5;6,7‐dibenzo‐3,8,12‐tri­oxa[8.3.1]­propellane has been determined previously by spectroscopic methods. The crystal structure of the di­spiro product, 2,7,12,17‐tetraoxa‐3,4:5,6:13,14:15,16‐tetrabenzodi­spiro[8.1.8.0]­nona­decane, C31H26O4, revealed that the conformations of the nine‐membered heterocyclic rings are due to interactions between the π‐electron system of the bi­phenyl moiety and the lone electron pairs of the ether O atoms, the repulsion of the lone electron pairs of atoms O1⃛O2 and O3⃛O4, and steric interactions between H atoms in ortho positions. The conformations have C1 symmetry and can be described approximately as twist‐boat.  相似文献   

12.
The title compounds, namely di­methyl (2RS)‐2,3‐di­phenyl‐1,3‐thia­zolidine‐5‐spiro‐2′‐adam­antane‐4,4‐di­carboxyl­ate methanol sol­vate, C28H31NO4S·0.275CH4O, and di­methyl (4RS)‐3,4‐di­phenyl‐1,3‐thia­zolidine‐5‐spiro‐9′‐(9′H‐fluorene)‐2,2‐di­car­box­ylate, C31H25NO4S, were obtained from dipolar [2+3]‐cyclo­additions of an azo­methine yl­ide with adamantane­thione and thio­fluorenone, respectively. The structures show that the choice of thio­ketone affects the regioselectivity of the cyclo­addition. The asymmetric unit of the former structure contains two mol­ecules of the thia­zolidine derivative plus a site for a partial occupancy (55%) methanol mol­ecule. O—H⋯O and C—H⋯O interactions link two of each of these entities into closed centrosymmetric hexamers. The five‐membered ring in each structure has an envelope conformation.  相似文献   

13.
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   

14.
Novel spiro[pyrazolo[3,4‐b]pyridine‐4,3′‐indoline] derivatives were prepared by the three‐component reaction of isatins 3‐methyl‐1‐phenyl‐1H‐pyrazol‐5‐amine and Meldrum's acid in the presence of a catalytic amount of melamine trisulfonic acid. This protocol provides a simple one‐step procedure with the advantages of easy work‐up, mild reaction conditions and environmentally benign.  相似文献   

15.
The reaction of homophthalic anhydride and N‐(1‐methyl‐1H‐pyrrol‐2‐yl‐methylidene)‐benzylamine in boiling benzene afforded as a main product the expected substituted trans‐1,2,3,4‐tetrahydroisoquinoline‐4‐carboxylic acid 5 . The carboxylic group of 5 was transformed in four steps into cyclic amino‐methyl groups yielding numerous new tetrahydroisoquinolinones 11a‐j incorporating a given fragment of pharmacological interest. Reduction of 11a‐j was studied.  相似文献   

16.
A group of 1,2‐diphenyl‐3,5‐dioxopyrazolidines possessing a methylsulfonyl ( 11 ) or sulfonamide ( 15 ) substituent at the para position of the N1‐phenyl ring, in conjunction with a hydrogen, methyl or fluoro sub‐stituent at the para position of the N2‐phenyl ring, and a C‐4 n‐butyl, methyl or spiro‐cyclopropyl substituent were synthesized for evaluation as potential cyclooxygenase‐2 (COX‐2) selective inhibitor antiinflammatory agents. The title compounds 11 and 15 were synthesized using a four‐step and a three‐step reaction sequence, respectively. Thus, the acetic acid promoted condensation of a nitrosobenzene 5 with an aniline derivative ( 6, 12 ) gave the corresponding azobenzene product ( 8, 13 ) which was reduced with zinc dust in the presence of ammonium chloride to yield the corresponding hydrazobenzene ( 9, 14 ). Base‐catalyzed condensation of 9 and 14 with a malonyl dichloride ( 10 ) afforded the target 3,5‐dioxopyrazolidine product ( 11,15 ). 4‐n‐Butyl‐1‐(4‐methylsulfonylphenyl)‐2‐phenyl‐3,5‐dioxopyrazolidine ( 11a ) was a selective COX‐1 inhibitor (COX‐1 IC50 = 8.48 μM). In contrast, 4‐n‐butyl‐1‐(4‐methylsulfonylphenyl)‐2‐(4‐tolyl)‐3,5‐dioxopyrazolidine ( 11b , COX‐2 IC50 = 11.45 μM) and 4‐n‐butyl‐1‐(4‐methylsulfonylphenyl)‐2‐(4‐fluorophenyl)‐3,5‐dioxopyrazoli‐dine ( 11c , COX‐2 IC50 = 9.86 μM) were about 46‐fold and 20‐fold less selective COX‐2 inhibitors respectively, relative to the reference drug celecoxib.  相似文献   

17.
Methyl (2E,4R)‐4‐hydroxydec‐2‐enoate, methyl (2E,4S)‐4‐hydroxydec‐2‐enoate, and ethyl (±)‐(2E)‐4‐hydroxy[4‐2H]dec‐2‐enoate were chemically synthesized and incubated in the yeast Saccharomyces cerevisiae. Initial C‐chain elongation of these substrates to C12 and, to a lesser extent, C14 fatty acids was observed, followed by γ‐decanolactone formation. Metabolic conversion of methyl (2E,4R)‐4‐hydroxydec‐2‐enoate and methyl (2E,4S)‐4‐hydroxydec‐2‐enoate both led to (4R)‐γ‐decanolactone with >99% ee and 80% ee, respectively. Biotransformation of ethyl (±)‐(2E)‐4‐hydroxy(4‐2H)dec‐2‐enoate yielded (4R)‐γ‐[2H]decanolactone with 61% of the 2H label maintained and in 90% ee indicating a stereoinversion pathway. Electron‐impact mass spectrometry analysis (Fig. 4) of 4‐hydroxydecanoic acid indicated a partial C(4)→C(2) 2H shift. The formation of erythro‐3,4‐dihydroxydecanoic acid and erythro‐3‐hydroxy‐γ‐decanolactone from methyl (2E,4S)‐4‐hydroxydec‐2‐enoate supports a net inversion to (4R)‐γ‐decanolactone via 4‐oxodecanoic acid. As postulated in a previous work, (2E,4S)‐4‐hydroxydec‐2‐enoic acid was shown to be a key intermediate during (4R)‐γ‐decanolactone formation via degradation of (3S,4S)‐dihydroxy fatty acids and precursors by Saccharomyces cerevisiae.  相似文献   

18.
Aldol condensation of 5,6‐dihydro‐4H‐pyrrolo[3,2,1‐ij ]quinoline‐1,2‐dione with aryl methyl ketones generates 3‐(aroylmethylidene)oxindoles, which react with hydrazine to generate tricyclic spiro[3H‐pyrazole‐3,3′‐oxindoles].  相似文献   

19.
In this study, (5α,7α)‐4,5‐epoxy‐3,6‐dimethoxy‐17‐methyl‐6,14‐ethenomorphinan‐7‐carboxylic acid hydrazide ( 5 ) was synthesized by the condensation of methyl (5α,7α)‐4,5‐epoxy‐3,6‐dimethoxy‐17‐methyl‐6,14‐ethenomorphinan‐7‐carboxylate ( 4 ) with NH2NH2⋅H2O. The (5α,7α)‐4,5‐epoxy‐3,6‐dimethoxy‐17‐methyl‐6,14‐ethenomorphinan‐7‐carboxylic acid 2‐[(arylamino)carbonyl]hydrazides 6a – 6q were prepared by the reaction of 5 with corresponding substituted aryl isocyanates, and the N‐{5‐[(5α,7α)‐4,5‐epoxy‐3,6‐dimethoxy‐17‐methyl‐6,14‐ethenomorphinan‐7‐yl]‐1,3,4‐oxadiazol‐2‐yl}arenamines 7a – 7q were obtained via the cyclization reaction of 6a – 6q in the presence of POCl3. The synthesized compounds have a rigid morphine structure, including the 6,14‐endo‐etheno bridge and the 5‐(arylamino)‐1,3,4‐oxadiazol‐2‐yl residue at C(7) adopting the (S)‐configuration (7α). The structures of the compounds were confirmed by high‐resolution mass spectrometry (HR‐MS) and various spectroscopic methods such as FT‐IR, 1H‐NMR, 13C‐NMR, APT, and 2D‐NMR (HETCOR, COSY, INADEQUATE).  相似文献   

20.
The resolution of 1‐i‐butyl‐3‐methyl‐3‐phospholene 1‐oxide was studied applying TADDOL [(−)‐(4R,5R)‐4,5‐bis(diphenylhydroxymethyl)‐2,2‐dimethyldioxolane], spiro‐TADDOL [(−)‐(2R,3R)‐α,α,α′,α′‐tetraphenyl‐1,4‐dioxaspiro[4.5]decan‐2,3‐dimethanol], or the acidic and neutral Ca2+ salts of (−)‐O,O′‐dibenzoyl‐ and (−)‐O,O′‐di‐p‐toluoyl‐(2R,3R)‐tartaric acid as the resolving agent. The absolute configuration of the P‐asymmetric center was determined by circular dichroism spectroscopy and related quantum chemical calculations. In one instance, the single crystal of the diastereomeric complex incorporating i‐butyl‐3‐phospholene oxide and spiro‐TADDOL was subjected to X‐ray analysis, which suggested a feasible hypothesis for the efficiency of the resolution process under discussion that may be an example for the “solvent‐inhibited” resolution.  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