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.
The 1H and 13C NMR spectra of the E and Z isomers of 2-, 3-, and 4-benzoylpyridine oximes and their ethers were analyzed thoroughly, and the 1H-13C spin-spin coupling constants (SSCC) were determined. It was established that the magnitude of the effect for the quaternary carbon atoms in the E and Z isomers depends on the site of substitution in the pyridine ring. It was assumed that the intermolecular hydrogen bond is stronger in the E form than in the Z form. The existence of the Z isomer of 2-benzoylpyridine oxime in deuterochloroform with an intramolecular hydrogen bond was proved.Translated from Khimiya Geterotsiklicheskikh Soedinenii, No. 2, pp. 203–208, February, 1990.  相似文献   

2.
Regioselective syntheses of novel 2‐(phosphoryl)methylidenethiazolidine‐4‐ones 3a–c, 5 by the condensation of phosphoryl acetic acid thioamides 2a–c or substituted thioanilide 4 with dimethyl acetylenedicarboxylate are described. N3‐unsubstituted thiazolidine‐4‐ones 3a–c were obtained as E,Z‐isomers, while N3‐phenyl substituted heterocycle 5 was formed as Z,Z‐isomer. The structures of thiazolidin‐4‐ones 3a ‐E,Z and 5 ‐Z,Z are characterized by crystal structure determination. According to B3Pw91/6‐31G* calculations, the isomers observed in crystals are thermodynamically preferable. In solutions, phosphorylated thiazolidines undergo isomerization (relative to C2 carbon atom of the heterocycle) proceeded by either imine–enamine (N3‐unsubstituted compounds 3a–c ) or push–pull mechanisms (N3‐substituted compound 5 ). © 2005 Wiley Periodicals, Inc. Heteroatom Chem 16:159–222, 2005; Published online in Wiley InterScience ( www.interscience.wiley.com ). DOI 10.1002/hc.20084  相似文献   

3.
TEMPO‐Mediated oxidation of hydroxylamines (=hydroxyamines) and alkoxyamines to the corresponding oxime derivatives is reported (TEMPO=2,2,6,6‐tetramethylpiperidin‐1‐yloxy radical; Scheme 2). These environmentally benign oxidations proceed in good to excellent yields (Table 1). For alkoxyamines, oxidation to the corresponding oxime ethers can be performed by using dioxygen as a terminal oxidant in the presence of 5–10 mol‐% of TEMPO or 4‐substituted derivatives thereof as a catalyst (Scheme 3 and Table 2). Importantly, benzyl bromides can directly be transformed to oxime ethers via in situ alkoxyamine formation by a nucleophilic substitution followed by TEMPO‐mediated oxidation (Scheme 4 and Table 3).  相似文献   

4.
The McLafferty rearrangement is an extensively studied fragmentation reaction for the odd‐electron positive ions from a diverse range of functional groups and molecules. Here, we present experimental and theoretical results of 12 model compounds that were synthesized and investigated by GC‐TOF MS and density functional theory calculations. These compounds consisted of three main groups: carbonyls, oximes and silyl oxime ethers. In all electron ionization mass spectra, the fragment ions that could be attributed to the occurrence of a McLafferty rearrangement were observed. For t‐butyldimethylsilyl oxime ethers with oxygen in a β‐position, the McLafferty rearrangement was accompanied by loss of the t‐butyl radical. The various mass spectra showed that the McLafferty rearrangement is relatively enhanced compared with other primary fragmentation reactions by the following factors: oxime versus carbonyl, oxygen versus methylene at the β‐position and ketone versus aldehyde. Calculations predict that the stepwise mechanism is favored over the concerted mechanism for all but one compound. For carbonyl compounds, C–C bond breaking was the rate‐determining step. However, for both the oximes and t‐butyldimethylsilyl oxime ethers with oxygen at the β‐position, the hydrogen transfer step was rate limiting, whereas with a CH2 group at the β‐position, the C–C bond breaking was again rate determining. n‐Propoxy‐acetaldehyde, bearing an oxygen atom at the β‐position, is the only case that was predicted to proceed through a concerted mechanism. The synthesized oximes exist as both the (E)‐ and (Z)‐isomers, and these were separable by GC. In the mass spectra of the two isomers, fragment ions that were generated by the McLafferty rearrangement were observed. Finally, fragment ions corresponding to the McLafferty reverse charge rearrangement were observed for all compounds at varying relative ion intensities compared with the conventional McLafferty rearrangement. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

5.
《Tetrahedron》1987,43(17):4051-4056
The Intramolecular nitrone cycloadditions to Z and E chiral allyl ethers afford annulated isoxazolidines with good to excellent stereocontrol in favour of the C-5/C-5' anti isomers. The relative stereochemistry at the stereocenters in C-3/C-4 depends on the length of the chain connecting dipole and dipolarophile.  相似文献   

6.
李玉玲a  b  顾大公a  徐小平a  纪顺俊  a 《中国化学》2009,27(8):1558-1562
本文研究了羟基化合物对两种类型N-对甲苯磺酰基氮杂环丙烷的开环反应。在功能性离子液体[hmim]HSO4存在条件下,氮杂环丙烷与醇反应,以中等到高的产率和非常高的区域选择性得到对应的β-胺基醚。并且离子液体[hmim]HSO4可以循环使用。  相似文献   

7.
The syntheses of some novel carboacyclic nucleosides, 17a – 17o , containing oxiconazole‐like scaffolds, are described (Schemes 13). In this series of carboacyclic nucleosides, pyrimidine as well as purine and other imidazole derivatives were employed as an imidazole successor in oxiconazole. These compounds could be prepared in good yields by using two different strategies (Schemes 1 and 2). Due to Scheme 1, the N‐coupling of nucleobases with 2‐bromoacetophenones was attained for 18a – 18e , and their subsequent oximation affording 19a – 19e and finally O‐alkylation with diverse alkylating sources resulted in the products 17a – 17g, 17n , and 17o . In Scheme 2, use of 2‐bromoacetophenone oximes 20 , followed by N‐coupling of nucleobases, provided 19f – 19j whose final O‐alkylation produced 17h – 17m (Scheme 2). For the rational interpretation of the dominant formation of (E)‐oxime ethers rather than (Z)‐oxime isomers, PM3 semiempirical quantum‐mechanic calculations were discussed and the calculations indicated a lower heat of formation for (E)‐isomers.  相似文献   

8.
Ketone oxime O-vinyl ethers having alkyl or phenyl radicals react with trifluoroacetic anhydride in ether in the presence of pyridine, yielding 43-54% of the corresponding ketone oxime O-(trans-4,4,4-trifluoro-3-oxo-1-butenyl) ethers with high stereoselectivity.  相似文献   

9.
In this article, we describe the characteristic 15N chemical shifts of isatin oxime ethers and their isomer nitrone. These oxime ethers and nitrones are the alkylation reaction products of isatin oximes. In our study, the 15N chemical shifts observed in these oxime ethers were in the 402–408 (or 22–28) ppm range, although those for their corresponding nitrone series were in the 280–320 (or ?100 to ?60) ppm range. This remarkable difference in 15N NMR chemical shift values could potentially be used to determine the Oversus N‐alkylation of oximes, even when only one isomer is available. In this paper, the differences in 15N NMR chemical shifts serve as the basis for a discussion about how to distinguish both regioisomers derived from the oximes alkylation. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

10.
A series of novel menthone oxime ethers were synthesized in three steps starting from (–)-menthol. Analysis of the 13C NMR chemical shift differences between α carbons of oxime derivatives (O-alkyl oximes) provides a convenient and reliable means of assigning oxime stereochemistry. It has been found that carbons syn to the oxime are shifted more upfield than carbons anti to the oxime moiety. Significant E products were obtained.  相似文献   

11.
The gas-phase rearrangement of (1Z, 2′E)-, (1Z, 2′Z)-, (1E, 2′E)-, and (1E, 2′Z)- propenyl but-2′-enyl ether (Z, E)-, (Z, Z)-, (E, E)-, and (E, Z)-1) into erythro- and threo-2, 3-dimethyl-pent-4-en-al (erythro- and threo-2) was investigated over a temperature range from 142,5° to 190,0° at 20–35 Torr (for kinetic data and activation parameters see table 2). All four stereoisomeric ethers 1 rearrange preferentially via a chair-like transition state C into the aldehydes 2 (ΔΔG (160°) = 2,5–2,7 kcal/mol for B – C (B = boat-like transition state). The relative rates (krel) for (Z, Z)-1, (Z, E)-1, (E,Z)-1, and (E,E)-1 at 160° are 1,0, 2,9, 4,3 and 9,0 respectively (see table 5). Taking into account the relative enthalpies of the ethers 1 and the steric interaction in the C transition state of the ethers 1 (see table 6), krel values can be estimated. They are in good agreement with those observed (see table 5).  相似文献   

12.
The E and Z geometric isomers of a stable silene (tBu2MeSi)(tBuMe2Si)Si=CH(1‐Ad) ( 1 ) were synthesized and characterized spectroscopically. The thermal Z to E isomerization of 1 was studied both experimentally and computationally using DFT methods. The measured activation parameters for the 1Z ? 1E isomerization are: Ea=24.4 kcal mol?1, ΔH=23.7 kcal mol?1, ΔS=?13.2 e.u. Based on comparison of the experimental and DFT calculated (at BP86‐D3BJ/def2‐TZVP(‐f)//BP86‐D3BJ/def2‐TZVP(‐f)) activation parameters, the Z?E isomerization of 1 proceeds through an unusual (unprecedented for alkenes) migration–rotation–migration mechanism (via a silylene intermediate), rather than through the classic rotation mechanism common for alkenes.  相似文献   

13.
The cis- and trans-propenyl alkyl ethers were polymerized by a homogeneous catalyst [BF3·O(C2H5)2] and a heterogeneous catalyst [Al2(SO4)3–H2SO4 complex]. Methyl, ethyl, isopropyl, n-butyl and tert-butyl propenyl ethers were used as monomers. The steric structure of the polymers formed depended on the geometric structures of monomer and the polymerization conditions. In polymerizations with BF3·O(C2H5)2 at ?78°C., trans isomers produced crystalline polymers, but cis isomers formed amorphous ones except for tert-butyl propenyl ether. On the other hand, highly crystalline polymers were formed from cis isomers, but not from the trans isomers in the polymerization by Al2(SO4)3–H2SO4 complex at 0°C. The x-ray diffraction patterns of the crystalline polymers obtained from the trans isomers were different from those produced from the cis isomers, except for poly(methyl propenyl ether). The reaction mechanism was discussed briefly on these basis of these results.  相似文献   

14.
Telaprevir is a potent, selective, peptidomimetic inhibitor of the hepatitis C virus (HCV) NS3‐4A serine protease. it is used for the treatment of HCV infection in combination with peginterferon alfa and ribavirin. In the present work, the E–Z isomerization process of telaprevir in solution was revealed by online HPLC–DAD (diode array detector)–MS, variable‐temperature and variable‐gradient experiments. The molecular geometry information of the two isomers was established by molecular mechanics calculations, and good correlation between the two isomers' UV–vis spectra and their molecular geometry information was also discovered. In addition, it was revealed by molecular docking that the two isomers have different affinities to HCV NS3?4A protease, and the Z isomer, the minor form of telaprevir in solution, is the more effective inhibitor of HCV NS3?4A protease. The investigation can provide more structure information about telaprevir in solution and in the binding process of HCV NS3?4A protease.  相似文献   

15.
Reactions of perfluoropropylene and its oligomers with acetone oxime in the presence of a base afford perfluoroalkyl and/or perfluoroalkenyl ethers of acetone oxime. When heated to 100 °C, the 3-perfluoro-2-methyl-2-pentenyl ether of acetone oxime (3) is quantitatively converted to 4-hydroxy-3-methyl-5,5-bistrifluoromethyl-4-pentafluoroethyl-1-pyrroline (4), the structure of which was established by X-ray diffraction analysis. A convenient one-stage synthesis of perfluoro-3-isopropyl-4-methyl-3-penten-2-one (7) was proposed.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 6, pp. 1068–1072, June, 1994.  相似文献   

16.
The reactivity of glycosylidene carbenes derived from pivaloylated vs. benzylated diazirines 1 and 2 towards enol ethers have been examined. The pivaloylated 1 led to higher yields of spirocyclopropanes than the benzylated 2. Among the enol ethers tested, dihydrofuran 6 proved most reactive, yielding 71–72% of the spiro-linked tetrahydrofuran 7 , while the benzylated diazirine 2 afforded only 33% of the analogue 8 (Scheme 1 ). Other enol ethers proved much less reactive. The addition of 1 and 2 to the dihydropyran 10 and the 2, 3-dihydro-5-methyl-furan 15 gave low yields of single cyclopropanes (→ 12 , 14 , and 16 ), and the glycals 17 and 18 , and (E)-1-methoxy-oct-1-ene ( 23 ) did not react. The main products of these reactions were the azines (Z, Z)- 11 and (Z, Z)/( E, E)- 13. Similarly, 1 and 2 reacted poorly with (Z)-1-methoxyoct-1-ene ( 24 ), leading to cyclopropanes 25 / 26 / 27 and 28 / 29 / 30 / 31 (Scheme 2). Main products were again the azines (Z, Z)- 11 and (Z, Z)/(E, E)- 13 . The structure of 70 and 25 was established by X-ray analysis (Figs. 1 and 2). The mechanism of addition of glycosylidene carbenes to enol ethers is discussed, AMI Calculations indicate that the LUMOcarbene/HOMOalkoxyalkene interaction is dominant at the beginning of the reaction, while the transition states are characterized by a dominant interaction of the doubly occupied, sp2-hybridized orbital of the carbene with the LUMO of the enol ether. The relative reactivity of the carbenes towards either the enol ethers or the diazirines determine type and yields of the products.  相似文献   

17.
Twelve novel compounds of 2‐acetyl‐1H‐benzimidazole oxime‐ethers and 2‐acetyl‐6‐chloro‐1H‐benzimidazole oxime‐ethers were synthesized with o‐phenylenediamine (or 4‐chloro‐o‐phenylenediamine), 2‐hydroxypropyl acid, alkoxy (or benzyloxy) amines hydrochloride as starting materials. The structures of the target compounds were characterized by IR, 1H NMR spectra and elemental analyses. The in vitro fungicidal activities against Botrytis cinerea Pers and Alternaria alternata were also evaluated by mycelium growth rate method. The results indicate that the compounds 3b , 3c , 3f , 3g and 3h exhibit good activities against Botrytis cinerea Pers, while 3b and 3f possess excellent activities against Alternaria alternate, and their fungicidal activities are all higher than that of carbendazim.  相似文献   

18.
In contrast to oxodipyrromethenes and bilirubin, benzalpyrrolinones (H, P-OCH3, p-Cl, p-N(CH3)2 and o-CH3) and α-pyridalpyrrolinones appear not to undergo dye-sensitized photo-oxygenation. They do, however, undergo an unsensitized E ? Z photoisomerization reminiscent of stilbene photoisomerization, and the photostationary state varies with substituent. Intramolecular H-bonding is implicated in the α-pyridalpyrrolinone isomerization. In each case, the Z isomers are the therrnodynamically more stable ones, but the corresponding E isomers have been isolated and characterized following photoirradialion.  相似文献   

19.
20, 21-Aziridine Steroids: Reaction of Derivatives of the Oximes of 5-Pregnen-20-one, 9β, 10α-5-Pregnen-20-one and 9β, 10α-5,7-Pregnadiene-20-one with Lithium Aluminium Hydride, and of 3β-Hydroxy-5-pregnen-20-one Oxime with Grignard Reagents. Reduction of 3β-hydroxy-5-pregnen-20-one oxime ( 2 ) with LiAlH4 in tetrahydrofuran yielded 20α-amino-5-pregnen-3β-ol ( 1 ), 20β-amino-5-pregnen-3β-ol ( 3 ), 20β, 21-imino-5-pregnen-3β-ol ( 6 ) and 20β, 21-imino-5-pregnen-3β-ol ( 9 ). The aziridines 6 and 9 were separated via the acetyl derivatives 7 and 10 . The reaction of 6 and 9 with CS2 gave 5-(3β-hydroxy-5-androsten-17β-yl)-thiazolidine-2-thione ( 8 ). Treatment of the 20-oximes 12 and 15 of the corresponding 9β,10α(retro)-pregnane derivatives with LiAlH4 gave the aziridines 13 and 16 , respectively. Their deamination led to the diene 14 and triene 17 , respectively. Reduction of isobutyl methyl ketone-oxime with LiAlH4 in tetrahydrofuran yielded 2-amino-4-methyl-pentane ( 19 ) as main product, 1, 2-imino-4-methyl-pentane ( 22 ) as second product and the epimeric 2,3-imino-4-methyl-pentanes 20 and 21 as minor products. – 3β-Hydroxy-5-pregnen-20-one oxime ( 2 ) was transformed by methylmagnesium iodide in toluene to 20α, 21-imino-20-methyl-5-pregnen-3β-ol ( 23 ) and 20β, 21-imino-20-methyl-5-pregnen-3β-ol ( 26 ). Acetylation of these aziridines was accompanied by elimination reactions leading to 3β-acetoxy-20-methylidene-21-N-acetylamino-5-pregnene ( 30 ) and 3β-acetoxy-20-methyl-21-N-acetylamino-5,17-pregnadiene ( 32 ). The reaction of oxime 2 with ethylmagnesium bromide in toluene gave 20α, 21-imino-20-ethyl-5-pregnen-3β-ol ( 24 ) and 20α,21-imino-20-ethyl-5-pregnen-3β-ol ( 27 ). Acetylation of 24 and 27 led to 3β-acetoxy-20-ethylidene-21-N-acetylamino-5-pregnene ( 31 ), 3β-acetoxy-20-ethyl-21-N-acetylamino-5,17-pregnadiene 33 and 3β, 20-diacetoxy-20-ethyl-21-N-acetylamino-5-pregnene ( 37 ). With phenylmagnesium bromide in toluene the oxime 2 was transformed to 20β, 21-imino-20-phenyl-5-pregnen-3β-ol ( 25 ) and 20β,21-imino-20-phenyl-5-pregnen-3β-ol ( 28 ). Acetylation of 25 and 28 yielded 3β-acetoxy-20-phenyl-21-N-acetylamino-5, 17-pregnadiene ( 34 ) and 3β,20-diacetoxy-20-phenyl-21-N-acetylamino-5-pregnene ( 39 ). LiAlH4-reduction of 39 gave 3β, 20-dihydroxy-20-phenyl-21-N-ethylamino-5-pregnene ( 41 ). – The 20, 21-aziridines are stable to LiAlH4. Consequently they are no intermediates in the formation of the 20-amino derivatives obtained from the oxime 2 .  相似文献   

20.
《Tetrahedron: Asymmetry》2003,14(11):1463-1466
Oxime ethers of acetophenone, isopropyl methyl ketone, and tert-butyl methyl ketone were reduced to the corresponding hydroxylamine ethers of 45–94% ee with borane–oxazaborolidine 1 derived from (−)-norephedrine. A one-pot reduction of acetophenone oxime with 1 to 1-phenylethylhydroxylamine of 87% ee is described. The reduction of 6-methyl-2,3,4,5-tetrahydropyridine and N-methylimines of the above mentioned ketones with borane-B-methyloxazaborolidine adduct 2, derived from (−)-diphenylprolinol, gave the corresponding amines of 40–74% ee.  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