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.
Treatment of the chloro-nitro-ribofuranose 7 with KPO(OMe)2 gave the O-amino phosphate 8 (5 %) and the nitrile 9 (62 %). Compound 9 was also obtained by the reaction of 8 with KPO(OMe)2, and its structure was established by X-ray analysis. Treatment of the chloro-nitro-mannofuranose 10 , the bromo-nitro-ribofuranose 14 , or the bromo-nitro-mannofuranose 16 , respectively, with the K or Na salt of HPO(OMe)2 lead also to O-amino phosphates and nitriles. The (1-C-nitroglycosyl)phosphonate 22 was obtained (21 %) together with the nitrile 21 (51 %) from the chloro-nitro-mannofuranose 10 and KPO(OEt)2. The reaction of the 1-C-nitroglycosyl sulfone 25 (NO2-group endo) with KPO(OEt)2 gave the (1-C-nitroglycosyl)phosphonate 22 (61%) and the nitrile 21 (11 %), whilst the anomeric sulfone 26 (NO2-group exo) gave 22 (15 %) and 21 (58 %). In the presence of [18] crown-6, a mixture of the anomers 25 and 26 gave the (1-C-nitroglycosyl)phosphonate 22 in 67 % yield together with 21 (13 %). These findings are rationalized as the result of a competition between a nucleophilic attack of the dialkyl-phosphite anions on the NO2-group leading ultimately to the nitrile 21 and a single-electron transfer reaction leading to the (1-C-nitroglycosyl)phosphonate 22 .  相似文献   

2.
The reaction of benzimidazo[1,2-b]isoquinolin-11(5H)-one with activated olefins has been studied. The derivatives of 3,10-dioxo-3H,10H-benzimidazo[1,2,3-ij]benzo[c][1,8]naphthyridine formed are the result of an initial Michael reaction at C(6) followed by intramolecular heterocyclization.  相似文献   

3.
An improved synthesis of trivinylaluminum (V3Al) is described. The proton magnetic resonance (PMR) spectrum of V3Al was recorded and analyzed. A new vinylation method involving the use of V3Al as the vinylating agent has been developed, and the vinylation of organic halides by V3Al was studied at ?30, ?50 and ?70°C. Primary alkyl chlorides, such as methyl and methylene chloride, do not react with V3Al and were used as solvents. Secondary chlorides such as 2-chloropropane also do not react. t-Butyl chloride gives rise to t-butylethylene (70–98%), depending on reaction conditions, and the allylic chlorides, 3-chloro-1-butene, and 3-chloro-3-methyl-1-butene, yield the expected vinylated products and their isomers (~90%). Allyl and benzyl chloride do not react under the conditions tried. The reaction between V3Al and the ditertiary dichloride 2,6-dichloro-2,6-dimethylheptane yields several isomeric C13H24 and C11H20 hydrocarbons; however, surprisingly, C9H16 does not form. The C13 hydrocarbons arise by divinylation at the termini of the dichloride, while the C11 hydrocarbons are formed by vinylation at one and proton elimination at the other terminus of the dichloride. The presence of unsaturated C13H24 and C11H20 isomers is most likely due to proton induced isomerization. These results are explained by a proximity effect involving vinylation at one end of the dichloride by V3Al followed by rapid reaction of the second chlorine (mostly) by V2AlCl generated in situ during the first vinylation in the proximity of the chloride. At the other chlorine terminus V2AlCl causes either a second vinylation (leading to C13 hydrocarbons) or a proton elimination (leading to C11 hydrocarbons). The absence of C9H16 among the reaction products indicates that V3Al exclusively effects vinylation. The RCl + V3Al ← RV + V2AlCl reaction may be regarded as a model for initiation followed by immediate termination in cationic olefin polymerization, a process leading to vinyl-ended polymers.  相似文献   

4.
In order to study the effects of substituents on the chain transfer reaction to cumenes, the polymerization of methyl methacrylate in a series of nuclear-substituted cumenes with α,α′-azobisisobutyronitrile as initiator was carried out at 60°C. and the chain transfer constants C were determined. The C values obtained for all substituted cumenes were greater than that (C0) for unsubstituted cumene, regardless of the electronattracting or -repelling nature of the substituents. Hence the plot of log (C/C0) against the Hammett σ constants gave no linear relationship. When the plots were made by the modified Hammett equation including resonance effect of the substituents: log (C/C) = ρσ + γER, however, a straight line with ρ = 0.03 and γ = 0.9 was obtained. From these results, it may be concluded that the effect of the substitutents on the chain transfer reactivities of cumenes toward attack of a poly(methyl methacrylate) radical is attributable mainly to the resonance contribution in the transition state. These results are also compared with those for polystyryl radical reported previously and discussed.  相似文献   

5.
The reaction of a series of β-methoxyvinyl trifluoromethyl ketones [CF3COC(R2)?C(OMe)R1, where R1 = Me, -(CH2)3-C3, -CH2)4-C3, Ph and R2 = H, Me, -(CH2)3-C4, -(CH2)4-C4] with N-methylhydroxylamine is reported. The regiochemistry of the reaction are explained by MO calculation data.  相似文献   

6.
We report a study of the conditions of the phosphorylation reaction for the preparation of aromatic polyamides using the Higashi reaction medium. For poly(p-phenylene terephthalamide) (PPD-T), the optimum conditions are: reaction temperature, 115°C; monomer concentration, C = 0.083 mol/L; and ratio of triphenyl phosphite (TPP) to monomer, 2.0. These optimum conditions produce PPD-T having ηinh = 6.2 dL/g. At temperatures of 120°C and above PPD-T precipitates from the reaction mixture, leading to lower molecular weights. At lower temperatures the reaction mixture gels, and the gel time decreases with increasing reaction temperature. However, polycondensation continues in the gel state. Monomer concentrations C = 0.10 mol/L and above produce precipitation and yield polyamides of lower molecular weight. For the preparation of poly(p-benzamide) (PBA), the optimum ratio of TPP to monomer is 0.6 for either p- aminobenzoic acid or N-4-(4′-aminobenzamido)benzoic acid. In the former case the inherent viscosity of polymer prepared at 115°C showed little dependence upon the concentration of the monomer. The highest value, ηinh = 1.8 dL/g, was obtained with C = 0.40 mol/L and a TPP/monomer ratio of 0.6. However, for the same TPP/monomer ratio, the monomer containing a preformed amide linkage, N-4-(4′-aminobenzamido)benzoic acid, gave PBA with ηinh = 4.6 dL/g when the monomer concentration is 0.33 mol/L. This is the highest value reported for PBA using the phosphorylation reaction. In A?A + B?B polycondensation, examples in which one of the monomers contained one or two preformed amide linkages produced polyamides having ηinh = 7.8 and 8.9 dL/g, respectively.  相似文献   

7.
Starting from the readily available, optically active (4R)-4-hydroxy-2,2,6-trimethylcyclohexanone ( 1 ), a new technical synthesis of (3R,3′R)-zeaxanthin is described. According to a 2(C9 + C6) + C10 = C40 construction scheme, the ketone 1 was first transformed with (E)-3-methylpent-2-en-4-yn-1-ol ( 5 ) into a C15-intermediate which, by a three-step sequence, could be converted into the known olefinic C15-Wittig salt 4 . Optimized conditions for the final Wittig reaction of 4 with the C10-dialdehyde 3 are discussed. Based on 1 , the overall yield of the entire technical process is ca. 40%.  相似文献   

8.
The NaNH2 catalyzed one stage reaction between phenylacetic acid dialkylamides and cinnamic acid methyl ester or dialkylamides was studied under various conditions. Conditions were found for easy preparation of each of the both possible diastereomeric derivatives of 2,3-diphenylglutaric acid. It was proved that catalytic amounts of NaNH2 take part in the reaction. It is assumed that the observederythro/threo equilibrium ratios are determined by an isomerization via two different carbanions (at C2 and C4) of the reaction products.Part III:J. Stefanovsky andL. Viteva, Comun. Dept. Chem. Bulg. Acad. Sci.4, 159 (1971).  相似文献   

9.
Phenol, 4-methoxyphenol, 4-nitrophenol, methyl orsellinate ( 1 ), and 2,6-di(tert-butyl)-4-methylphenol (BHT; 2 ) have been glycosylated by thermal reaction (20–60°) with various glycosylidene-derived diazirines. 4-Methoxyphenol reacted with the D-glucosylidene-derived diazirine 3 to give O-glucosides ( 4 and 5 , 69%, 3:1) and C-glucosides ( 6 and 7 , 16%, 1:1). Similarly, phenol yielded O-glucosides ( 10 and 11 , 70%, 4:1) and C-glucosides ( 12 and 13 , 13%, 1:1). 4-Nitrophenol gave only O-glycosides, 3 leading to 14 and 15 (75%, 3:2; Scheme 1), and the D-galactosylidene-derived diazirine 17 to 22 and 23 (52% (from 16 ), 65:35; Scheme 2). The reaction of phenol with 17 yielded 58% (from 16 ) of the O-galactosides 18 and 19 (4:1) and 14% of the C-galactosides 20 and 21 (1:1). From the D-mannosylidene-derived diazirine 25 , we predominantly obtained the α-D-configurated 26 (38 % from 24 ). These results are interpreted by assuming that an intermediate (presumably a glycosylidene carbene) first deprotonates the phenol to generate an ion pair which combines to give O- and - with electron-rich phenolates - also C-glycosides. A competition experiment of 3 with 4-nitro- and 4-methoxyphenol gave the products from the former ( 14 and 15 ) and the latter phenol ( 4-7 ) in almost equal amounts. Differences in the kinetic acidity of OH groups, however, may form the basis of a regioselective glycosidation, as evidenced by the reaction of 3 with methyl orsellinate ( 1 ) yielding exclusively the 4-O-monoglycosylated products 27 and 28 (78%, 85:15), although diglycosidation is possible ( 27 → 31 and 32 ; 67%, 4:3; Scheme 3). Steric hindrance does not affect this type of glycosidation; 3 reacted with the hindered BHT ( 2 ) to afford 33 and 34 (81 %, 4:1). The predominant formation of 1,2-trans -configurated O-aryl glycosides is rationalized by a neighbouring-group participation of the 2-benzyloxy group.  相似文献   

10.
The sphingolipids 1a , b and 2a , b which play important roles in epidermal barrier function, were synthesized by N-acylation of C18-sphingosine 3 and 1-O-glucosylated C18-sphingosine 6 , respectively, with ω-acyloxy-substituted fatty acids 4 and 5 (Scheme 1). These fatty acids were obtained from ω-hydroxy-substituted fatty acids 8 and 9 by esterification with linoleic acid ( 7 ). The C34-fatty acid 8 was prepared as follows: C25-Compound 18 was obtained by means of a Wittig reaction of C13-aldehyde 13 with C12-phosphonium salt 15 or of C12-aldehyde 24 with C13-phosphonium salt 21 , respectively, and subseqent hydrogenation and O-deprotection (Scheme 2). Alternatively, 8 was prepared via 30 by copper-catalyzed coupling of C13-alkyl halide 19 with the Grignard reagent derived from C12-alkyl bromide 14 (Scheme 2). Oxidation of 18 to aldehyde 39 and Wittig reaction with C9-phosphonium salt 41 furnished the desired ω-hydroxy-substituted fatty acid 8 , after O-deprotection (Scheme 3). Similarly, Wittig reaction of C11-phosphonium salt 22 with C12-aldehyde 24 furnished C23-aldehyde 40 , after hydrogenation, O-deprotection, and oxidation; Wittig reaction with compound 41 and subsequent deprotection afforded the desired C32-fatty and 9 (Scheme 3). an alternative strategy furnished compound 8 by a coupling reaction of alkyne 53 with ω-bromo-substitued fatty acid 52 , obtained from compounds 24 and 47 by Wittig reaction, hydrogenation, and introduction of bromide (Scheme 4). Hydrogenation (Lindlar's catalyst) of the resulting C34-alkyne 54 and deprotection furnished 8 .  相似文献   

11.
Poly(vinyl chloride) pendant with polysulfide (PS–PVC) having various degrees of substitution, various S substituents, and various numbers of atoms in the sulfur chain has been synthesized by the reaction of poly(vinyl chloride) with a thiol, sulfur, and triethylamine in dimethylformamide at 30°C for 0.4–5 hr. The photocrosslinking reaction has been investigated under ultraviolet irradiation at 250–450 mμ. The photocrosslinking reaction of PS–PVC is influenced by the degree of substitution, the nature of the S substituent, and the number of atoms in the sulfur chain. The degree of photocrosslinking r increased in the order, n-C4H9? < n-C8H17? < C6H5CH2? < i-C3H7? < t-C4H9? . On the photocrosslinking of PS–PVC having two different S substituents, r increases in the similar order for aliphatic substituents and in the order NO2C6H4? < ClC6H4? < C6H5CH2? < CH3C6H4? < t-C4H9C6H4? < C6H5? for the aromatic substituents. Further, r increases markedly with the increase of sulfur chain number for all PS–PVC. The chemical structure of the crosslinks and the crosslinking mechanism are discussed on the basis of the results.  相似文献   

12.
A redetermination of the disproportionation/combination ratio for n–C3F7 and C2H5 radicals gives a value of Δ(n–C3F7, C2H5) = 0.13 ± 0.01, independent of the temperature. The radicals were produced by the photolysis of n–C3F7COC2H5. The previous determinations of this ratio are discussed and are found to be largely incorrect. The values for Δ(CF3, C2H5) and Δ(C2F5, C2H5) are also re-evaluated, and the recommended values are 0.10 ± 0.02 and 0.12 ± 0.02, respectively. Systems involving perfluoroalkyl and ethyl radicals are complicated due to rapid perfluororadical addition to the ethylene formed in the disproportionation process. The extent of this reaction, and its consequences, are discussed and evaluated. The role of the propionyl (C2H5CO) radical in the room temperature photolysis is also assessed. However, it is found that the Δ values determined by the intercept method used in this work are not affected by the secondary reactions that occur. It is concluded that high cross-combination ratios are general to perfluoroalkyl-alkyl radical interactions. For C3F7 and C2H5 radicals the ratio is 2.7–2.8. Above 100°C ratios exceed 3 due to secondary reactions.  相似文献   

13.
The title complex, tetra‐μ‐acetato‐O:O′‐bis{[μ‐1,4‐bis(2‐­pyridyl­oxy)­phenyl­ene‐N,C2:N′,C6]dipalladium(II)} bis­(tri­chloro­methane) dihydrate, [Pd4(C16H10N2O2)2(C2H3O2)4]·2CHCl3·2H2O, the product of the reaction of 1,4‐bis(2‐pyridyl­oxy)­benzene with palladium acetate, is shown to be a tetranuclear, rather than a polymeric, species. It crystallizes about a centre of inversion and has two doubly cyclo­palladated ligands bridged by four acetate groups. The cyclo­palladated ligand is far from planar in the complex and has the central benzene rings π‐stacked. The chelate rings exist in shallow boat conformations.  相似文献   

14.
Total Synthesis of Natural α-Tocopherol Two independent syntheses of (S)-6-hydroxy-2,5,7,8-tetramethylchroman-2-yl-methanol ( 8b ), (Scheme 6 resp. 9) as optically active chroman moiety for the preparation of natural vitamin E via (S)-6-acetoxy-2,5,7,8-tetramethylchroman-2-carboaldehyde ( 2a ) (Scheme 1) and a corresponding side chain are described. Both reaction sequences use trimethyl-hydroquinone as starting material; one approach employs an optically active C4 unit ( 10a ) (Schemes 5 and 6) to introduce the required configuration at C(2), the other uses an optically active C5-synthon ( 11a ) (Schemes 8 and 9) to build the optically active chroman unit. The correct configuration and optical purity of the chroman synthesized is established by correlation with optically pure material of known configuration from which natural vitamin E had already been derived [2].  相似文献   

15.
The synthesis of methyl (4R, 8R,)-10-bromo-8-methyl-4-(1,3,6-trioxaheptane)-2-deceneoate ( 5 ), a synthon for the construction of the macrocyclic moieties of the cytochalasins A ( 1), B. (2) F (3) and desoxaphomin ( 4 ) is described. (S)-Glutamic acid ( 6 ) was transformed to the C5-epoxide 10 and 3-methylglutaric acid ( 11 ) to the C5-bromide 15 . Coupling of both 10 and 15 by a CuI-catalyzed Grignard reaction gave the decanol 16 in very high yield. The latter was transformed by several steps to synthon 5 .  相似文献   

16.
The reaction C2H5 + O2 → C2H5O2 in glassy methanol-d4 and the H-atom abstraction by CH3, C2H5, and n-C4H9 radicals in C2H5OH + C2D5OH and CD3CH2OH + C2D5OH glassy mixtures have been studied by electron spin resonance. The analysis of the dependence of the reaction rates on the concentration of O2 (oxidation) and C2H5OH, CD3CH2OH (H-atom abstraction) has shown that the √t law is not conditioned by the existence of regions characterized by different rate constants.  相似文献   

17.
18.
In connection with the accelerating effect of poly(N-vinylpyrrolidone) (PVP) on a Williamson reaction, PhONa + NC4H9Br → PhO-(n-C4H9) + NaBr, the ionic dissociation of sodium phenoxide (PhONa) was studied by means of conductance measurements and ultraviolet spectroscopy. Results suggest that the degree of dissociation of PhONa increased with the amount of PVP as well as N-methylpyrrolidone (NMP), the monomeric analog, the effect of PVP being much larger than that of NMP at the same concentration. It is assumed that free phenolate anion produced by solvation of sodium cation with NMP or the pyrrolidone residues of PVP plays an important role in the acceleration of the reaction and that the higher reaction rates in PVP solution are due to the greater dissociation of PhONa.  相似文献   

19.
The reactivity of the free aluminylene [N]-Al ( 1 ) ([N]=1,8-bis(3,5-di-tert-butylphenyl)-3,6-di-tert-butylcarbazolyl) towards boron Lewis acids is investigated. A facile oxidative addition reaction of 1 with Ph2BOBPh2 furnishes an exceedingly scarce example of the free alumaborane [N]-Al(BPh2)(OBPh2) ( 2 ) with an Al−B electron-sharing bond. By contrast, complexation of 1 with B(C6F5)3 and HB(C6F5)2 gives rise to the corresponding Lewis adducts [N]-Al→B(C6F5)3 ( 3 ) and [N]-Al→BH(C6F5)2 ( 4 ), respectively, with an Al→B dative bond. Crystallization of 4 in Et2O produces the adduct [N]-Al(Et2O)→BH(C6F5)2 ( 5 ). Quantum chemical calculations are carried out to understand the formation of 2 as well as the bonding situation of 3 and 5 .  相似文献   

20.
In this communication, an unprecedented interception of CnF2n+1(O)SO. radical with a copper-based carbene has been established. Distinguished by wide substrate scopes and mild reaction conditions, this novel radical–carbene coupling reaction (RCC reaction) provides a fundamentally different and mechanistically interesting strategy for the synthesis of perfluoroalkanesulfinate esters.  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