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 mechanism of a recently discovered intramolecular Heck‐type coupling of oximes with aryl halides (Angew. Chem. Int. Ed. 2007 , 46, 6325) was systematically studied by using density functional methods enhanced with a polarized continuum solvation model. The overall catalytic cycle of the reaction was found to consist of four steps: oxidative addition, migratory insertion, β‐H elimination, and catalyst regeneration, whereas an alternative base‐promoted C? H activation pathway was determined to be less favorable. Migratory insertion was found to be the rate determining step in the catalytic cycle. The apparent activation barrier of migratory insertion of the (E)‐oxime was +20.5 kcal mol?1, whereas the barrier of (Z)‐oxime was as high as +32.7 kcal mol?1. However, (Z)‐oxime could isomerize to form the more active (E)‐oxime with the assistance of K2CO3, so that both the (E)‐ and (Z)‐oxime substrates could be transformed to the desired product. Our calculations also indicated that the Z product was predominant in the equilibrium of the isomerization of the imine double bond, which constituted the reason for the good Z‐selectivity observed for the reaction. Furthermore, we examined the difference between the intermolecular Heck‐type reactions of imines and of olefins. It was found that in the intermolecular Heck‐type coupling of imines, the apparent activation barrier of migratory insertion was as high as +35 kcal mol?1, which should be the main obstacle of the reaction. The analysis also revealed the main problem for the intermolecular Heck‐type reactions of imines, which was that the breaking of a C?N π bond was much more difficult than the breaking of a C?C π bond. After systematic examination of a series of substituted imines, (Z)‐N‐amino imine and N‐acetyl imine were found to have relatively low barriers of migratory insertion, so that they might be possible substrates for intermolecular Heck‐type coupling.  相似文献   

2.
Refluxing the oxime ( 1 ) of 2-methyl-2H-1,2-benzothiazin-4(3H)one 1,1-dioxide with tri-fluoroacetic acid or with boron trifluoride in acetic acid gives the corresponding N-acyl derivative ( 2 or 3) of 4-amino-2-methyl-2H-1,2-benzothiazin-3(4H)one 1,1-dioxide. This transformation appears to be related to the acid catalyzed conversion of α-tetralone oxime to α-naphthylamine.  相似文献   

3.
The bulk polymerization and copolymerization of N-(1,1′-dimethyl-3-oxobutyl) acrylamide oxime have been studied. Polymerization of diacetone acrylamide oxime was carried out with different initiating systems. The rate of polymerization of diacetone acrylamide oxime with azoisobutyronitrile as the initiating system was much higher than with peroxides. However, in the case of perester initiating systems (t-butyl perbenzoate and t-butyl per ethyl-2-hexanoate), cobalt salt promoted the polymerization rate markedly. Diacetone acrylamide oxime readily formed copolymers with a variety of comonomers (crosslinking agents and reactive diluents). Gel permeation chromatography has shown a higher reactivity of diacetone acrylamide oxime with trimethylol propane trimethacrylate as crosslinking agent and N-vinyl-pyrrolidone as reactive diluent. Therefore, the dynamic mechanical analyses presented an increase in Tg with trimethylolpropane trimethacrylate and N-vinyl-pyrrolidone as comonomers. The terpolymer formed with diacetone acrylamide oxime, trinethylolpropane trimethacrylate, and N-vinyl-pyrrolidone exhibited interesting mechanical properties and high temperature behavior.  相似文献   

4.
Bis­[μ‐1‐(3‐pyridyl)­ethanone oxime‐κ2N:N′]­bis­[nitrato­sil­ver(I)], [Ag2(NO3)2(C7H8N2O)2], crystallizes as a centrosymmetric binuclear macrocylic complex containing silver(I) ions bridged by the organic 1‐(3‐pyridyl)­ethanone oxime ligand. The ligand coordinates via the pyridine and the oxime N atoms. A similar metal–ligand arrangement was found in the copper(I) complex catena‐poly­[[bis­[μ‐1‐(3‐pyridyl)­ethano­ne oxime‐κ2N:N′]­dicopper(I)]‐di‐μ‐iodo], [Cu2I2(C7H8N2O)2]n, but here the centrosymmetric macrocycles are connected by double anion bridges, resulting in the formation of a one‐dimensional coordination polymer.  相似文献   

5.
The crystal structure of the title complex, tetrakis­[μ-6-amino-3-methyl-4-aza­hex-3-en-2-one oximato(1–)-κ4N,N′,N′′:O]tetracopper(II) tetraperchlorate 0.6-hydrate, [Cu4(C6H12N3O)4](ClO4)4·0.6H2O, shows the cation to be an oximate-bridged tetramer composed of four 6-amino-3-methyl-4-aza­hex-3-en-2-one oxime ligands and four copper(II) ions and to have crystallographically imposed symmetry. Each CuII atom is four-coordinated by the three N atoms of one oxime ligand and by the O atom of another oxime ligand in a distorted square-planar geometry.  相似文献   

6.
The chalcone as (E)-1-([1,1′-biphenyl]-4-yl)-3-(3,4-dimethoxyphenyl)prop-2-en-1-one ( 3 ) was reacted with various active methylene compounds via Michael addition reaction under different conditions. In one hand, chalcone 3 reacted with isatin and glycine in one pot reaction via 1,3-dipolar cycloaddition reaction. On the other hand, chalcone 3 was also reacted with different N-nucleophiles via direct addition on the carbonyl group to award cyclic and/or acyclic products. Meanwhile, the reaction of chalcone 3 with S-benzylthiuronium chloride afforded the thio-Michael addition product. Chalcone 3 and 10 novel synthesized compounds were screened against two cell lines (HepG2 and MCF-7). Among of them, thiosemicarbazone 16 , oxime 14 and pyrimidine-2(1H)-thione 19 derivatives revealed an excellent activity against both cell lines (IC50 values = 6.79-12.91 μM), whereas thiosemicarbazone 16 (6.79 ± 0.5 and 7.58 ± 0.6 μM) showed the highest activity.  相似文献   

7.
1-Methyl­indole-3-carbox­aldehyde oxime, C10H10N2O, (I),and (E)-5-methoxy-1-methyl­indole-3-carbox­aldehyde oxime, C11H12N2O2, (II), were ex­amined structurally to ascertain the geometry of the hydroxy­imino function relative to the indole core. Oxime (I) exhibits cis geometry and there are two mol­ecules in the asymmetric unit. In contrast, oxime (II) exhibits trans geometry and has four mol­ecules in the asymmetric unit, with the geometry of the 5-methoxy group in one mol­ecule differing from that in the other three. Both crystal structures are maintained by hydrogen bonding with no π-stacking of the indole moiety present.  相似文献   

8.
Isobaric product ions cannot be differentiated by exact mass determinations, although in some cases deuterium labeling can provide useful structural information for identifying isobaric ions. Proposed fragmentation pathways of fentanyl were investigated by electrospray ionization ion trap mass spectrometry coupled with deuterium labeling experiments and spectra of regiospecific deuterium labeled analogs. The major product ion of fentanyl under tandem mass spectrometry (MS/MS) conditions (m/z 188) was accounted for by a neutral loss of N‐phenylpropanamide. 1‐(2‐Phenylethyl)‐1,2,3,6‐tetrahydropyridine (1) was proposed as the structure of the product ion. However, further fragmentation (MS3) of the fentanyl m/z 188 ion gave product ions that were different from the product ion in the MS/MS fragmentation of synthesized 1, suggesting that the m/z 188 product ion from fentanyl includes an isobaric structure different from the structure of 1. MS/MS fragmentation of fentanyl in deuterium oxide moved one of the isobars to 1 Da higher mass, and left the other isobar unchanged in mass. Multistage mass spectral data from deuterium‐labeled proposed isobaric structures provided support for two fragmentation pathways. The results illustrate the utility of multistage mass spectrometry and deuterium labeling in structural assignment of isobaric product ions. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

9.
Beckmann reaction products of 11-dihomodriman-8α-ol-12-one oxime with Ac2O in pyridine, 86% H3PO4, p-TsCl in pyridine, and PCl5 in ether were investigated. It has been found that the major product from treatment of the oxime with Ac2O is the oxime acetate. Reaction of the oxime with 86% H3PO4 gave (1S,2S,4aS,8aS)-2,5,5,8a-tetramethyldecahydro-1H-naphtho[1, 2][5, 6]-3-methyl-4,5-dihydro[1, 2, 6]oxazine; with p-TsCl, (1S,2S,4aS,8aS)-2,5,5,8a-tetramethyldecahydro-1H-naphtho[1, 2][5, 6]-2-methyl-4,5dihydro[1, 3, 6]oxazine; with PCl5, a mixture of products containing 11-acetylamino- and 11-methylaminooxodrimenes that were isomeric at the double bond, norambreinolide, and a 1,3,6-oxazine.  相似文献   

10.
Condensation of tetra-O-pivaloyl-α-d-glucopy-ranosyl bromide (1) with three heterocyclic oximes: 3-hydroxyiminoquinuclidine (2), 4-hydroxyiminomethyl-pyridine (3) and N-methyl-2-hydroxyiminomethylimidazole (4) leads to the β-N-glucoconjugates. Conjugates 6 and 7 were synthesized using aromatic compounds 3 and 4 as the starting material. They were obtained in two isomeric forms (E and Z) due to the restricted rotation around the oxime's double bond. The presence of E and Z isomers was proved by comparison of NMR spectra with calculated GIAO/DFT NMR spectra on B3LYP/6-31G(d) level of theory and by X-ray structural analysis of starting oxime reagents. Isomery was not observed in the quinuclidinium glucoconjugate 5.  相似文献   

11.
The diastereoselectivity of the addition of NH3 and MeNH2 to glyconolactone oxime sulfonates and the structures of the resulting N‐unsubstituted and N‐methylated glycosylidene diaziridines were The 15N‐labelled glucono‐ and galactono‐1,5‐lactone oxime mesylates 1* and 9* add NH3 mostly axially (>3 : 1; Scheme 4), while the 15N‐labelled mannono‐1,5‐lactone oxime sulfonate 19* adds NH3 mostly equatorially (9 : 1; Scheme 7). The 15N‐labelled mannono‐1,4‐lactone oxime sulfonate 30* adds NH3 mostly from the exo side (>4 : 1; Scheme 9). The configuration of the N‐methylated pyranosylidene diaziridines 17, 18, 28 , and 29 suggests that MeNH2 adds to 1, 9, 19 , and 23 mostly to exclusively from the equatorial direction (>7 : 3; Schemes 5 and 8). The mannono‐1,4‐lactone oxime sulfonate 30 adds MeNH2 mostly from the exo side (85 : 15; Scheme 10), while the ribo analogue 37 adds MeNH2 mostly from the endo side (4 : 1; Scheme 10). Analysis of the preferred and of the reactive conformers of the tetrahedral intermediates suggests that the addition of the amine to lactone oxime sulfonates is kinetically controlled. The diastereoselectivity of the diaziridine formation is rationalized as the result of the competing influences of intramolecular H‐bonding during addition of the amines, steric interactions (addition of MeNH2), and the kinetic anomeric effect. The diaziridines obtained from 2,3,5‐tri‐O‐benzyl‐D ‐ribono‐ and ‐D ‐arabinono‐1,4‐lactone oxime methanesulfonate ( 42 and 48 ; Scheme 11) decomposed readily to mixtures of 1,4‐dihydro‐1,2,4,5‐tetrazines, pentono‐1,4‐lactones, and pentonamides. The N‐unsubstituted gluco‐ and galactopyranosylidene diaziridines 2, 4, 6, 8 , and 10 are mixtures of two trans‐substituted isomers ( S / R ca. 19 : 1, Scheme 2). The main, (S,S)‐configured isomers S are stabilised by a weak intramolecular H‐bond from the pseudoaxial NH to RO? C(2). The diaziridines 12 , derived from GlcNAc, cannot form such a H‐bond; the (R,R)‐isomer dominates ( R / S 85 : 15; Scheme 3). The 2,3‐di‐O‐benzyl‐D ‐mannopyranosylidene diaziridines 20 and 22 adopt a 4C1 conformation, which does not allow an intramolecular H‐bond; they are nearly 1 : 1 mixtures of R and S diastereoisomers, whereas the OH5 conformation of the 2,3:5,6‐di‐O‐isopropylidene‐D ‐mannopyranosylidene diaziridines 24 is compatible with a weak H‐bond from the equatorial NH to O? C(2); the (R,R)‐isomer is favoured ( R / S ≥7 : 3; Scheme 6). The mannofuranosylidene diaziridine 31 completely prefers the (R,R)‐configuration (Scheme 9).  相似文献   

12.
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   

13.
14.
The Beckmann rearrangement of either syn or anti 3,4-diphenyl-cyclohexenone oxime 2a,b in polyphosphoric acid produces only one of the two possible isomeric unsaturated caprolactams 1 . Under neutral conditions, only the syn oxime tosylate 9b rearranges to lactam 1 , the anti oxime tosylate 9a remains unchanged. These results support earlier reports that alkyl migration is preferred over vinyl migration in the Beckmann rearrangement of unsaturated cyclic ketoximes. Structure proof of the lactam was made using deuterium exchange and HMQC nmr experiments.  相似文献   

15.
Three new oxime‐based palladacycles, namely [Pd{C,N‐C6H4{C(Me)?NOH}‐2}(dppm)]ClO4 ( 1 ), [Pd2{C,N‐C6H4{C(Me)?NOH}‐2}2(dppe)2(μ‐dppe)](ClO4)2 ( 2 ) and [Pd{C,N‐C6H4{C(Me)?NOH}‐2}(dppmS2)]ClO4 ( 3 ), were synthesized by the reaction of dinuclear oxime complex [Pd{C,N‐C6H4{C(Me)?NOH}‐2}(μ‐Cl)]2 with different diphosphine ligands (dppm, dppe and dppmS2). The synthesized complexes were characterized using Fourier transform infrared, 31P NMR, 1H NMR and 13C NMR spectroscopic methods and elemental analyses, and their molecular structures were elucidated using X‐ray crystallography. The structure of 2 is worthy of note as it is the first oxime palladacycle where there are both bridging (P–) and chelating (P^P) dppe ligands, giving rise to a dinuclear complex. The palladium atom is in a five‐coordinate, square pyramidal P3NC environment, while in 3 the palladium atom is in a distorted square planar environment, coordinated by the oxime ligand and a chelating (S^S) dppmS2 ligand. These complexes were employed as efficient catalysts for the Suzuki–Miyaura cross‐coupling reaction of several aryl bromides with phenylboronic acid. The in vitro cytotoxicity of the compounds was also evaluated against human tumour cell lines (HT29, A549 and HeLa) using the MTT assay method. The results indicate that the dinuclear complex 2 has greater catalytic and anticancer activity in comparison with the mononuclear complexes 1 and 3 .  相似文献   

16.
The configurational assignment of Z and E nitrogen derivatives () of 3,5,5-trimethyl-2-cyclohexen-1-one was made taking into consideration the through-space effects on oxime, O-methyloxime, dimethylhydrazone, tert-butylimine, N,N,N -trimethylhydrazonium iodide and oxime hydrochloride derivatives. The relationship between the magnitude of the chemical shifts of the α-protons and the dihedral angle formed by the α-C? H bond and the C?N? OH plane was interpreted in terms of the geometrical dependence of the electric field effect. For the different Y substituents, the change in chemical shift between the Z and the E configuration of the proton near the functional group was mainly dependent on the electric field effect.  相似文献   

17.
《Tetrahedron letters》2004,45(29):5563-5567
Reactions of (E)- and (Z)-ketoximes with trialkylphosphines and diphenyl disulfide (PhSSPh) have been compared to gain insight into the mechanisms involved and their potential applications. N-Sulfenylimine isomers and ketimines have been spectroscopically characterised. Both the E and Z isomers of erythromycin A oxime, when treated with Bu3P and PhSSPh (1:4:8 ratio), give the same N-phenylsulfenyl ketimine (of configuration E) as the major compound, whereas with Bu3P or Me3P and PySeSePy (1:8:4 ratio) afford the imine in good yield. Clarithromycin oxime behaves similarly.  相似文献   

18.
The known D -gluconhydroximo-1,5-lactam (= D -glucono-1,5-lactam oxime) 7a , its nitrogen isotopomers 7b and 7c , and the N-arylcarbamates 26–29 were synthesized from 2,3,4,6-tetra-O-benzyl-D -glucono-1,5-lactam ( 11a ) and its nitrogen isotopomer 11b to establish the controversial structure of 7a and to study the inhibition of β-glucosidases by the N-arylcarbamates 26–29 . Conversion of 11a with Lawesson's reagent yielded a mixture of the thionolactam 15a and its manno-configurated isomer 16a , which was transformed into a mixture of the benzylated hydroximo-lactam 13a and the manno-isomer 17a . Debenzylation (Na/NH3) and acetylation of this mixture led to the gluco-configurated pentaacetate 14a and the manno-isomer 18a . Treatment of 11a with Et3O·BF4 and then with H2NOH gave exclusively the benzylated D -gluconhydroximo-1,5-lactam (benzylated D-nojirilactam oxime) 13a , which was transformed into 14a . Deacetylation of 14a yielded the hydroximo-lactam 7a . The isotopomers 7b and 7c were obtained by analogous reaction sequences, using either 15NH3 or 15NH2OHHCl. To prepare the acetylated N-arylcarbamates 20–25 , 13a was debenzylated and acetylated (→ 14a ), followed by selective deacetylation to the tetraacetate 19a and treatment with the appropriate isocyanates. The structure of the 2-chlorophenyl carbamate 21 was established by X-ray analysis. Deacetylation of 20–23 led to the N-arylcarbamates 26–29 . The 15N-NMR spectra of 7b , 7c , and of their precursors 13b , 13c , 14b , and 14c , show that the C?N bond in all these lactam oximes is exocyclic as predicted from semiempirical and ab initio SCF-MO calculations on the structure of acetamide oxime and 5-pentanelactam oxime. According to these calculations, 5-pentanelactam oxime is a (Z)-configurated, flattened chair. X-ray analysis established the structure of D -glucono-1,5-lactam oxime ( 7a ) in the solid state, where it adopts a conformation between 4C1 and 4H3. In H2O, 7a is a flattened 4C1. The calculations also predict that protonation at the exocyclic N-atom strengthens the conjugation between the endocyclic N-atom and the hydroxyimino group, and leads to a half-chair conformation. This is evidenced by the chemical shift differences in the 15N-NMR spectra observed upon protonation of 7b and 7c . The hydroximolactam 7a and the N-arylcarbamates 26–29 are competitive inhibitors of the β-glucosidases from sweet almond (emulsin) and from Agrobacterium faecalis (= Abg), with KI values between 8 and 21·10?6M against emulsin (at pH 6.8) and between 0.15 and 1.2·10?6M against Abg (at pH 7.0).  相似文献   

19.
Rearrangement of Vinyl-Cyclopropane-Carbaminals Both (c-2, t-3-diphenyl-r-1-cyclopropyl)methylene-dipyrrolidine ( 4 ) and its (t-2, t-3)-isomer 10 underwent a thermal rearrangement to (E)-N-2-benzylidene-1-indanyl-pyrrolidine ( 5 ). Under the conditions of the rearrangement, 5 was partially converted into 2-benzyl-1-indanone ( 6 ) in a base catalysed reaction. The structures of 5 and 6 were derived from spectroscopic data and from degradation reactions.- N,N′-(t-2-Vinyl-r-1-cyclopropyl)methylene-dipyrrolidine ( 11 ) rearranged thermally to N-(2-methylidene-3-cyclopenten-1-yl)pyrrolidine ( 12 ), the structure of which was established from spectroscopic evidence and from a hydrogenation to N-(2-methylcyclopentyl)pyrrolidine ( 13 , cis/trans mixture 3:2). The aminal 4 was reduced with formic acid to give N-(c-2, t-3-diphenyl-r-1-cyclopropyl)methyl-pyrrolidine ( 14 ). If perdeuterio formic acid was used, the mixture product 14-d/14-d 2 was obtained which contained exactly one deuterium atom in its methylene group and about half a deuterium atom on C(1). This labeling pattern is mechanistically explained with the existence of a fast equilibrium between the iminium ion 19 and the enamine 18 , so that 18 and 19 are considered to be plausible reactive intermediates in the above mentioned thermal rearrangement. - Based on this, several mechanisms for the rearrangements 4 → 5, 10 → 5 and 11 → 12 were considered: A Pictet-Spengler- or Mannich-type reaction, which starts from the iminium ion 23 and is followed by a cyclopropylmethyl-homoallylic rearrangement and by deprotonation (path a, Scheme 5), was judged to be improbable because the postulated intermediates could lead more easily to other stable products than the observed ones. If the reaction is formulated as a [3,3]-sigmatropic shift occurring on exclusively the (E)-isomer 5 suggests a concerted process whose steric course is predominantly controlled by strain factors. Alternatively, the reaction could be formulated via a dipolar ( 27 ) or a diradical ( 26 ) species derived from the enamine 22 (paths c and d, Scheme 5); attempts to trap such species by a number of agents were unsuccessful. - The previously unknown aminals 10 and 11 were synthesized by standard methods.  相似文献   

20.
The N-phenylcarbamate 7 , derived from 2-acetamido-2-deoxy-D -gluconhydroximo-1,5-lactone ( 3 ) and the analogous N-phenylcarbamate 14 , derived from chitobionhydroximo-1,5-lactone ( 10 ) have been prepared as potential inhibitors of β-N-acetylglucosaminidases. The unambiguous synthesis of the hydroximo-1,5-lactone 3 involves oxidation of the oxime 1 , followed by deprotection with Na/NH3.  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