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.
The tetrazole 1 is a new β-glucosidase inhibitor (IC50=8·10?5 M , Emulsin), obtained (92%) by deprotection of 22 , the product of an intramolecular cycloaddition of the azidonitrile 20 . This azidonitrile was formed as an intermediate by treating the L -ido-bromide 14 or the L -ido-tosylate 19 with NaN3 at 110–120°. It was isolated in a separate experiment. The yield of 22 from 19 reached 70%; 21 was formed as by-product (10%). The bromide 14 (42%) and the iodide 15 (30–35%) were obtained from the nitrile 13 , together with the 2,5-anhydro-L -idononitrile 16, which was formed in ca. 35–45%. The tosylate 19 was obtained from 18 (97%). To obtain 18 , the nitrile 13 was oxidized according to Swern (→17, 92%) and then reduced (NaBH4, CeCl3), leading to 18 and 13 (92%, 18/13 93:7). Reduction of the tetrahydropyridotetrazole 22 with LiAlH4 afforded 83 % of the piperidine 23 , which was deprotected to (+)-1-deoxynojirimycin hydroacetate (2·AcOH, 86%) and further converted into the corresponding hydrochloride and into the free base 2 .  相似文献   

2.
Several 2‐alkylquinolizinium‐1‐olates 9 , i.e., heterobetaines, were prepared from ketone 11 , the latter being readily available either from pyridine‐2‐carbaldehyde via a Grignard reaction, followed by oxidation with MnO2, or from 2‐picolinic acid (=pyridine‐2‐carboxylic acid) via the corresponding Weinreb amide and subsequent Grignard reaction. Mesoionic heterobetaines such as quinolizinium derivatives have the potential to undergo cycloaddition reactions with double and triple bonds, e.g., 1,3‐dipolar cycloadditions or Diels? Alder reactions. We here report on the scope and limitations of cycloaddition reactions of 2‐alkylquinolizinium‐1‐olates 9 with electron‐poor acetylene derivatives. As main products of the reaction, 5‐oxopyrrolo[2,1,5‐de]quinolizines (=‘[2.3.3]cyclazin‐5‐ones’) 19 were formed via a regioselective [2+3] cycloaddition, and cyclohexadienone derivatives, formed via a Diels? Alder reaction, were obtained as side products. The structures of 2‐benzylquinolizinium‐1‐olate ( 9a ) and two ‘[2.3.3]cyclazin‐5‐ones’ 19i and 19l were established by X‐ray crystallography.  相似文献   

3.
Ethyl (1R, 2S)-2-hydroxycyclopentanecarboxylate and -cyclohexanecarboxylate ( 1a and 2a , respectively) obtained in 40 and 70% yield by reduction of 3-oxocyclopentanecarboxylate and cyclohexanecarboxylate, respectively (Scheme 2), with non-fermenting yeast, are converted to bicyclic dioxanone derivatives 3 and 4 with formaldehyde, isobutyraldehyde, and pivalaldehyde (Scheme 3). The Li-enolates of these dioxanones are alkylated (→ 5a – 5i , 5j , 6a – 6g ), hydroxyalkylated (→ 51, m, 6d, e ), acylated (→ 5k, 6c ) and phenylselenenylated (→ 7 – 9 ) with usually high yields and excellent diastereoselectivities (Scheme 3, Tables and 2). All the major isomers formed under kinetic control are shown to have cis-fused bicyclic structures. Oxidation of the seleno compounds 7–9 leads to α, β-unsaturated carbonyl derivatives 10 – 13 (Scheme 3) of which the products 12a – c with the C?C bond in the carbocyclic ring (exocyclic on the dioxanone ring) are most readily isolated (70–80% from the saturated precursors). Michael addition of Cu(I)-containing reagents to 12a – c and subsequent alkylations afford dioxanones 14a – i and 16a – d with trans-fused cyclohoxane ring (Scheme 4). All enolate alkylations are carried out in the presence of the cyclic urea DMPU as a cosolvent. The configuration of the products is established by NMR measurements and chemical correlation. Some of the products are converted to single isomers of monocyclic hydroxycyclopentane ( 17 – 19 ) and cyclohexane derivatives ( 20 – 23 ; Scheme 5). Possible uses of the described reactions for EPC synthesis are outlined. The observed steric course of the reactions is discussed and compared with that of analogous transformations of monocyclic and acyclic derivatives.  相似文献   

4.
Carbocycles from monosaccharides. III. Concerning the diastereoselective formation of cyclopentane derivatives. Transformations in the galactose series. The diastereoselectivity of the intramolecular nitrone-olefine cycloaddition of 1 , 3 and 4 (Scheme 1), yielding only 2 , 5 and 6 but none of the isomers 8 , 9 and 10 is explained by assuming a kinetic control and postulating that the relative activation energies of the two relevant transition states in the cyclization of e.g. 1 can be estimated from the conformers A and B , the latter being destabilized by a synperiplanar arrangement of the nitrone function and the 2-alkoxy-group (Scheme 2). It is further postulated, that this destabilization is responsible for the formation of (2,3)-trans configurated products. Since 2 , 5 and 6 are presumably thermodynamically more stable than 8 , 9 and 10 , a case was investigated, where the cycloaddition can either give thermodynamically less stable (2,3)-trans-product such as 12 or a thermodynamically more stable (2,3)-cis-product such as 13. 12 and 13 could both be formed from the aldehyde 25 via the nitrone 11 (Schemes 3 and 5). Treatment of the galactoside 16 first with Zn in aqueous butanol (forming among other products 25 and its 2-debenzyl-oxy-derivative) and then with N-Methyldroxylamine yielded the isoxazolidines 12 (72%), 13 (2%) and 27 (7%) (Schemes 4 and 6). Similarily, the anomeric silylated galactosides 17 and 23 gave 29 (78% from 17 , 77% from 23 ) and 27 (5% and 3%). Upon desilylation, 29 gave 32 , which was converted into 12 . The structure of the isoxazolidines was unambiguously deduced from their NMR. spectra and those of their derivatives 33 and 34 . Compound 32 was further transformed into its deoxyderivative 36 . The high diastereoselectivity of the cycloaddition restricts the number of diastereomeric, pentasubstituted cyclopentanes available by this method. However, cyclization of the 2-Hydroxy-aldehyde 37 (Scheme 8) gave the kinetically less favoured isomer 40 in a higher proportion, showing the differential influence of hydrogen-bonds on the relevant activation energies. Thermolysis of 32 gave 40 (79%) and 41 (11%). The structure of 41 was deduced from its NMR. spectra and those of its derivatives 42 and 43 . Thermolysis of 29 gave, after desilylation, 41 (42%), 40 (22%) and 32 (13%) and thermolysis of 6 lead to a 25 : 75 equilibrium with 44 (combined yield 90%). These transformations illustrate means leading to additional isomers and are in agreement with the proposed explanation of the diastereoselectivity in question.  相似文献   

5.
To obtain stable derivatives of α-unsubstituted pyrroles, the reaction of the test pyrrole 9 with a series of chalcones 14a – h were studied. Michael adducts 16b – h could be isolated. In order to synthesize coloured derivatives, the reaction of different pyrroles 9, 21, 23 , and 25 with diphenylpropynone 19 was investigated. In these cases, too, Michael-addition products were formed. The intense absorption band around 400 nm makes the identification of these derivatives easy.  相似文献   

6.
The regio‐ and stereoselectivity of cycloadditions of the nitrone 1a and the chiral, sugar‐derived nitrones 13a and 13b with 3‐(prop‐2‐enoyl)‐1,3‐oxazolidin‐2‐one ( 2 ) depends on the nature of the Lewis acid catalyst used. Addition of Lewis acid reverses the regioselectivity of the cycloaddition, and improves the anti‐diastereoselectivity in the case of chiral nitrones. The sterically favored isoxazolidin‐5‐yl‐substituted adducts 3, 4 , and 14 – 17 are produced as the major products in the absence of Lewis acid, while the electronically favored regioisomers with isoxazolidin‐4‐yl substituents ( 5, 6 , and 18 – 21 , respectively) are obtained as major products in the [Ti(OiPr)2Cl2] catalyzed reactions. The reactions of nitrone 13b with 2 in the presence of other Lewis acids such as ZnCl2, ZnBr2, ZnI2 and MgI2/I2 gave both regioisomeric pairs of the diastereoisomers, favoring the 4‐substituted congeners. The diastereoisomeric isoxazolidines 3a – 6a were reduced with NaBH4 in THF/H2O with subsequent desilylation to yield the separable diols 9 – 12 . Reduction of the diastereoisomeric isoxazolidines 19a and 18a afforded the chiral alcohols 23 and 22 , the latter of which was analyzed by X‐ray crystallography.  相似文献   

7.
Difluoro(iminio)methanides generated by the action of difluorocarbene on Schiff bases react with derivatives of maleic and fumaric acids, following the 1,3-dipolar cycloaddition pattern to give 2,2-difluoropyrrolidines which were detected by gas chromatography-mass spectrometry. The final products are stereo-isomeric substituted 2-pyrrolidinones formed by hydrolysis of difluoropyrrolidines and their dehydrofluorination products, 2-fluoro-4,5-dihydropyrroles. The observed stereoselectivity of the cycloaddition suggests Z configuration of intermediate ylide and both endo- and exo-approach to the dipolarophile in the transition state corresponding to cycloaddition.  相似文献   

8.
Carbocyclic compounds from monosaccharides. II. Transformations in the mannose series Upon treatment with Zn in refluxing aqeous ethanol or butanol, the anomeric mannopyranosides 11 and 14 yielded the aldehyde 15 which was subjected (via its N-methyl-nitrone) to an intramolecular nitrone-olefine cycloaddition leading diastereoselectively to 18 in a yield of 64% from 14 . Minor products of this reaction sequence were the compounds 19 , 20 and 21 . Similarily, the easily accessible unsaturated furanose 26 , upon treatment with N-methyl-hydroxylamine gave the isoxazolidines 27 (84%) and 28 (3%), thus showing that a free 4-hydroxy group is not detrimental to this intramolecular nitrone-olefin cycloaddition. The configuration of 18 , 27 and 28 (resp. 29 ) were determined by spectroscopic means; that of 27 was further proven by transformation into the compound 31 , of which an X-ray analysis was performed.  相似文献   

9.
The [3,3′(4H,4′H)‐bi‐2H‐1,3‐oxazine]‐4,4′‐diones 3a – 3i were obtained by [2+4] cycloaddition reactions of furan‐2,3‐diones 1a – 1c with aromatic aldazines 2a – 2d (Scheme 1). So, new derivatives of bi‐2H‐1,3‐oxazines and their hydrolysis products, 3,5‐diaryl‐1H‐pyrazoles 4a – 4c (Scheme 3), which are potential biologically active compounds, were synthesized for the first time.  相似文献   

10.
The reaction of different 5-amino-3-Q-1H-1,2,4-triazoles 1 with ethyl 2-cyano-3-ethoxyacrylate ( 5a ) and 2-cyano-3-ethoxyacrylonitrile ( 5b ) to yield either the a type 5-amino-, or the b type 7-amino-1,2,4-triazolo[1,5-a]-pyrimidine derivatives 6–10 was studied. The structure of compounds 6 and 9 was proved by their degradation to the corresponding derivatives 17a and 18a , respectively, through intermediates 11a, 12a, 13a, 14a, 15a and 16a , respectively. The structure of derivatives 7, 8 and 10 was proved on the basis of the analogy of their uv spectra with those of 6a and 9a , respectively. The isolation of the intermediates 19 and 20 helped to prove the mechanism of the reactions leading to the formation of 6a and 9a , respectively. In the reaction of the N-substituted 5-amino-1,2,4-triazoles with 5a the expected condensed ring products were not formed. Instead the aminoacrylates 22 and 24 were obtained. The “Z”-“E” isomeric structure of derivatives 19, 20, 22 and 24 was proved with the help of their pmr spectra. The “Z” isomeric structure of the thermodynamically stabile 22 was corroborated with the help of its proton coupled cmr spectra, too.  相似文献   

11.
Intramolecular 1,3-Dipolar Cycloadditions of Diaryl-nitrile-imines Generated from 2,5-Diaryl-tetrazoles Alkenyl-substituted diaryl-nitrile-imines – generated by photolysis or thermolysis of alkenyl-substituted 2,5-diaryl-tetrazoles - undergo a regioselective intramolecular [2 + 3] cycloaddition to yield new heterocyclic compounds, e. g. fused 2-pyrazolines. With alkinyl derivatives, the corresponding pyrazoles have been formed. UV evidence is given for the intermediate nitrile-imine at ? 190°. The latter can be trapped using an excess of carboxylic acid (UV evidence for a new intermediate at ? 120°). In this case, the corresponding rearranged addition product N′-acyl-N′-aryl-benzohydrazide is isolated in good yield.  相似文献   

12.
The reactions of azetidin-3-ones 6 – 10 , readily available from the amino acids L -alanine, L -phenylalanine, L -valine, L -lysine, and L -aspartic acid, via the corresponding diazo ketones, with nucleophilic reagents such as complex hydrides, Grignard compounds, an ester enolate, and a Wittig ylide give the expected products 11 – 19 in good yields and mostly in high diastereoselectivities. New amino-alcohol, γ-amino- and γ-amino-β-hydroxy-carboxylic-acid derivatives of known configurations are thus available.  相似文献   

13.
Photochemical behaviors of the pyrazinone derivatives 5,6,7,8-tetrahydroquinoxalin-2(1H)-ones 1a – c and 1,5,6,7,8,9-hexahydro-2H-cyclohepta[b]pyrazin-2-one 1d were investigated. Dye-sensitized photo-oxygenation of 1a-c gave the 1:1 adducts 5a – c of the corresponding 3,8a-epidioxy-3,5,6,7,8,8a-hexahydroquinoxalin-2(1H)-one 4 and H2O, whereas 1d gave 3,9a-epidioxy-1,3,5,6,7,8,9,9a-octahydro-2H-cyclohepta[b]pyrazin-2-one 4d (Scheme 2). The different kind of products was interpreted as being the result of the ring strain and steric hindrance of endoperoxides produced from 1a – d with singlet oxygen. Irradiation of 1a – b in the presence of alkenes gave tricyclic azetidine derivatives 9 by [2 + 2] cycloaddition of the C?N bond of 1 to the alkene.  相似文献   

14.
The photochemical synthesis of indole derivatives starting from the indoline-2-thiones 1 is described. Irradiation of indoline-2-thiones 1 in the presence of alkenes 3 gave 2-alkyl-3H-indoles 4 – 7 or 2-alkylindoles 8 – 22 through the ring cleavage of the intermediates, spirocyclic amino-thietanes, initially derived by [2 + 2] cycloaddition of the C?S bond of 1 and the C?C bond of 3 . Irradiation of 1 in the presence of trialkylamines 26 gave desulfurization products 27 – 32 and unexpected 3-alkylindoles 33 – 40 . N-Acylindoline-2-thiones 11 - p yielded the deacylated products, indoline-2-thiones 1a - b , and ethyl esters 43 through γ-H abstraction by the excited thioamide S-atom when irradiated in CDC13/EtOH or benzene/EtOH. Oxygen analogues 2a - d also underwent intramolecular H abstraction to give the indolin-2-ones 2e – f and ethyl esters 43 in a similar way.  相似文献   

15.
1,5-Dipolar Electrocyclization of Acyl-Substituted ‘Thiocarbonyl-ylides’ to 1,3-Oxathioles The reaction of α-diazoketones 15a, b with 4,4-disubstituted 1,3-thiazole-5(4H)-thiones 6 (Scheme 3), adamantanethione ( 17 ), 2,2,4,4-tetramethyl-3-thioxocyclobutanone ( 19 ; Scheme 4), and thiobenzophenone ( 22 ; Scheme 5), respectively, at 50–90° gave the corresponding 1,3-oxathiole derivatives as the sole products in high yields. This reaction opens a convenient access to this type of five-membered heterocycles. The structures of three of the products, namely 16c, 16f , and 20b , were established by X-ray crystallography. The key-step of the proposed reaction mechanism is a 1,5-dipolar electrocyclization of an acyl-substituted ‘thiocarbonyl-ylide’ (cf. Scheme 6). The analogous reaction of 15a, b with 9H-xanthen-9-thione ( 24a ) and 9H-thioxanthen-9-thione ( 24b ) yielded α,β-unsaturated ketones of type 25 (Scheme 5). The structures of 25a and 25c were also established by X-ray crystallography. The formation of 25 proceeds via a 1,3-dipolar electrocyclization to a thiirane intermediate (Scheme 6) and desulfurization. From the reaction of 15a with 24b in THF at 50°, the intermediate 26 (Scheme 5) was isolated. In the crude mixtures of the reactions of 15a with 17 and 19 , a minor product containing a CHO group was observed by IR and NMR spectroscopy. In the case of 19 , this side product could be isolated and was characterized by X-ray crystallography to be 21 (Scheme 4). It was shown that 21 is formed – in relatively low yield – from 20a . Formally, the transformation is an oxidative cleavage of the C?C bond, but the reaction mechanism is still not known.  相似文献   

16.
We report herein the interaction of diethylethoxymethylene malonate ( 1 ) with 2‐cyanomethylbenzothiazole ( 7 ) to give diethyl 2‐(2‐benzothiazole‐2‐(3H)‐ylidiene)‐2‐(cyano ethyl) malonate ( 8a ) in excellent yield. Ethyl 4‐cyano‐1‐oxo‐1H‐benzo[4,5]thiazolo[3,2‐a]pyridine‐2‐carboxylate (9) was synthesized from 8a and subjected to react with hydrazine hydrate to give its corresponding acid hydrazide 10 . Condensation of 10 with different acid anhydrides afforded the corresponding benzothiazolo pyridine carboxamide derivatives 11 – 15 . In addition, we report a simple synthesis of N′‐(benzo[d]thiazol‐2‐yl)‐2‐((4‐ayl)amino)acetohydrazide derivative ( 17 ), which then reacted with different amines to give the corresponding acetohydrazide derivatives 19a – c . Moreover, compound 17 reacted with some sulfonamide derivatives to give the corresponding sulfonamide derivatives 20 and 22a , b .The newly synthesized compounds were established their structures on the bases of their correct analytical and spectral data and evaluated their antimicrobial activity. It was found that compounds 22a , b displayed the highest antimicrobial activity against the tested organisms.  相似文献   

17.
Hydroxyalkylations of Cysteine through the Enolate of (2R,5R)-2(tert-Butyl)-1-aza-3-oxa-7-thiabicyclo[3.3.0]octan-4-one with Self-Reproduction of the Center of Chirality The heterobicyclic compound 1 specified in the title is readily prepared as a single stereoisomer from (R)-cysteine, formaldehyde, and pivalaldhyde. While it is not possible to generate the enolate 10 from 1 qunatitatively – due to β-elimination of thiolate (→6) – an in-situ addition to aromatic aldehydes such as benzaldehydes (→13–16) , pyrrol-, furan-, and thiophen-2-carbaldehydes (→17–19) , pyridine-3-carbaldehyde (→21) , as well as to other non-enolizable aldehydes like cinnamaldehyde (→22) , can be achieved in yields of ca. 50%. The adducts ( 8 and 9 ) of lithium diisopropylamide or t-butoxide to these aldehydes are acting, probably as bases for deprotonation and as in-situ sources of the electrophilic aldehyde species (cf. 11, 12 ). - Of the four possible diastereoisomeric products, one is usually formed with >90% selectivity (Table). It is assumed that the preferred stereochemical course of the reaction corresponds to that observed previously with the analogous proline-derived enolate (See 23,24 ). A chemical correlation with l-α-methyl-β-phenylserine (25) proves the relative configuration of the benzaldehyde adduct 13 . All hydroxyalkylated products (13–19, 21, 22) are obtained as crystalline, diastereoisomerically pure compounds and are fully characterized. – The benzaldehyde derivative 13 was used to exemplify the various possible transformations of these products to monocyclic or acyclic amino-acid derivatives such as the oxazolidionenes 26 and 29 (cleavage of the ring containing the S -atom), the thiazolidines 28 , 31 , and 32 (cleavage of the cyclic N,O-acetal) and the α-branched cysteine 27 and the phenylserines 25 and 30 (cleavage of both rings to give open-chain aminoacids).  相似文献   

18.
Glucosyl, cellobiosyl, cellotriosyl, cellotetraosyl, and cellooctaosyl residues were attached to naphthalene-1,8-diethanol ( 3 ) with the goal of preparing mimics of cellulose I. Among the templates that were considered, 1,8-diethynylnaphthalene ( 1 ) led to unstable products, and glycosidation of naphthalene-1,8-dimethanol ( 2 ) gave orthoesters that could not be rearranged to glycosides (Scheme 1). The conformation of 3 in the crystal and of its dimethyl ether 14 in solution was studied by X-ray analysis and force-field calculation (Figs. 1 – 3). Rotation around the Ar−CH2 and CH2−CH2 bonds of 14 is only weakly hindered and the O⋅⋅⋅O distance of crystalline 3 (6.01 Å) corresponds to the mean distance of the parallel chains of cellulose Iβ. The acetylated glycosyl bromides 18 and 19 were prepared by a new convergent synthesis (Scheme 2). Glycosylation of 3 by the glycosyl bromides 15 – 19 under established conditions of the Koenigs-Knorr reaction proved problematic, particularly on account of an acetyl transfer blocking one of the hydroxyethyl groups. Basic zinc carbonate, however, promoted glycosylation of 12 and 3 by the glycosyl bromides 15 – 19 and did not lead to transacetylation (Scheme 3). The mono- to tetrasaccharides 32 – 35 and 42 – 45 were isolated in yields of 56 – 82%, and the octasaccharides 36 and 46 in 32 and 16%, respectively. The mono- and disaccharides 32 , 33 , 42 , and 43 were deacetylated with NaOMe in MeOH. Aqueous NaOH was used for the tri-, tetra-, and octasaccharides 34 – 36 and 44 – 46 , as their partially deacetylated derivatives proved insoluble in MeOH. The fully deprotected saccharides 37 – 41 and 47 – 50 were isolated in over 90%, while the yield of the dioctaoside 51 was lower on account of its poor water solubility.  相似文献   

19.
Lewis acid or Brønsted acid catalyzed reactions of vinylidene cyclopropanes (VDCPs), 1 , with activated carbon–nitrogen, nitrogen–nitrogen, and iodine–nitrogen double‐bond‐containing compounds have been thoroughly investigated. We found that pyrrolidine and 1,2,3,4‐tetrahydroquinoline derivatives can be formed in good yields in the reactions of VDCPs 1 with ethyl (arylimino)acetates 2 by a [3+2] cycloaddition or intramolecular Friedel–Crafts reaction pathway. Based on these results, we found that activated carbon–nitrogen and nitrogen–nitrogen double‐bond‐containing compounds, such as N‐toluene‐4‐sulfonyl (N‐Ts) imines 5 and diisopropylazodicarboxylate ( 7 ), can also react with VDCPs 1 to give [3+2] cycloaddition products in moderate to good yields in the presence of a Lewis acid. When Ntert‐butoxycarbonyl aldimine 9 was used as the substrate, six‐membered cycloaddition products 10 and 11 were formed in moderate yields in the presence of a Brønsted acid, trifluoromethanesulfonic acid (TfOH). The reactions of VDCPs 1 with N‐Ts‐iminophenyliodinane ( 12 ) were also carried out in the presence of (CuOTf)2 ? C6H6 and it was found that nitrogen‐containing indene derivatives 13 were obtained, rather than the aziridination products. Plausible mechanisms for all of these transformations are discussed, based on the obtained results.  相似文献   

20.
Photoinduced Cycloadditions of 2,2-Dimethyl-3-phenyl-2H-azirine with Nitriles and ‘push-pull’ Olefines. Electron deficient nitriles of the type 5a–e in contrast to nonactivated nitriles undergo regiospecific [2+3]cycloadditions to benzonitrile isopropylide ( 2b ), which was generated in situ by irradiation of 2,2-dimethyl-3-phenyl-2H-azirine ( 1b ), to yield the 2H-imidazole derivatives 6a – e (Scheme 2). The structure of the photoproducts was mainly deduced from 13C-NMR. and mass spectrometry. Whereas normal olefins or enolethers do not react with 2b , push-pull olefins of the type 10a – d readily undergo the cycloaddition to give the 3-alkoxy-5,5-dimethyl-2-phenyl-1-pyrrolines 11a – d (Scheme 3 and 4). The structure of the photoproducts 11a – d indicates that the regiospecificity of the cycloaddition corresponds to that of acrylonitriles and acrylesters with 2b .  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