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 46 毫秒
1.
A new and convenient synthesis of 7‐(3‐chloropropoxy)‐4‐hydroxy‐6‐methoxyquinoline‐3‐carbonitrile, the key intermediate to bosutinib, is described on a hectogram scale. 5‐Bromo‐2‐methoxyphenol is adopted as the starting material via the simple chemical process including Friedel‐Crafts reaction, alkylation, bromination, cyano substitution, and so on to give the 3‐amino‐2‐(2‐bromobenzoyl)acrylonitrile compound 25 , which underwent key intramolecular cyclization at K2CO3/DMF condition; the title product was obtained in 36.9% yield over 7 steps and 98.71% purity (HPLC).  相似文献   

2.
4‐(Nitro, amino, acetylamino)‐2‐aminobenzoic acid were allowed to react with PPh3(SCN)2 and gave the crossholding 7‐nitro, 7‐acetylamino‐ and 7‐amino‐2‐thioxo quinazolin‐4‐ones respectively. The nature of the substituent at position 4 of the 2‐aminobenzoic acids has significant influence on the outcome of the cyclisation reaction with PPh3(SCN)2. Similarly, the nature of the substituent at position 7 of the 2‐substituted quinazolin‐4‐ones significantly affected the ease with which alkylation reactions could be performed. The alkylation selectivity of the 7‐ substiuted‐2‐thioxo quinazolin‐4‐ones was found to depend on the nature of the alkyl halide and the nature of the substituent at position 2.  相似文献   

3.
C9 fraction is the by‐products of catalytic reforming and ethylene cracker; it is considered as a kind of petroleum resin raw material. Recently, it was studied for the use as a gasoline additive to enhance the economic benefits. However, C9 fractions are getting higher in sulfur contents. As conventional hydrotreating technology leads to significant octane number loss and processing costs, the alkylation desulfurization process, which could reduce the sulfuric compounds to hydrogen sulfide by catalytic alkylation with olefins and distillation followed by, is a rather attractive way of reducing the sulfur of C9 fraction. In this paper, different kinds of thiophenic compounds, including 2‐ethylthiophen, 2,5‐dimethylthiophene, and 2‐n‐propylthiophene, were selected as the model compounds. Thiophenic compounds were studied first by the alkylation reaction over macroporous sulfonic resin Amberlyst 36, and the octane number of C9 fraction was measured. It was found that isoamylene and Amberlyst 36 resin had a significant effect on the alkylation desulfurization of thiophenic compounds with the conversion, reaching to above 99%. And the octane number of C9 fraction was increased by alkylation desulfurization over Amberlyst 36 resin. Moreover, the alkylation of thiophenic sulfurs could be described as a pseudo–first–order reaction as well as the reaction rate constant, and the activation energy of alkylation reactions was calculated.  相似文献   

4.
4‐Aryldihydropyrimidinones were obtained in excellent yields via solid‐phase Biginelli reaction of arylaldehydes, β‐ketoester and urea, using bismuth nitrate, immobilized on Al2O3 as catalyst. Subsequently, the products were converted into corresponding tetrahydropyrimidino[4,3‐a]isoquinolines (69–73% yield) by one‐pot regiospecific N‐alkylation‐annulation reaction using sodium hydride and dimethylaminopyridine. J. Heterocyclic Chem., (2011).  相似文献   

5.
A highly enantioselective Friedel–Crafts (F–C) alkylation of indoles and pyrrole with chalcone derivatives catalyzed by a chiral N,N′‐dioxide–Sc(OTf)3 complex has been developed that tolerates a wide range of substrates. The reaction proceeds in moderate to excellent yields and high enantioselectivities (85–92 % enantiomeric excess) using 2 mol % (for indole) or 0.5 mol % (for pyrrole) catalyst loading, which showed the potential value of the catalyst system. Meanwhile, a strong positive nonlinear effect was observed. On the basis of the experimental results and previous reports, a possible working model is proposed to explain the origin of the activation and asymmetric induction.  相似文献   

6.
The kinetics of propylene polymerization initiated by ansa‐metallocene diamide compound rac‐Me2Si(CMB)2Zr(NMe2)2 (rac‐1, CMB = 1‐C5H2‐2‐Me‐4‐tBu)/methylaluminoxane (MAO) catalyst were investigated. The formation of cationic active species has been studied by the sequential NMR‐scale reactions of rac‐1 with MAO. The rac‐1 is first transformed to rac‐Me2Si(CMB)2ZrMe2 (rac‐2) through the alkylation mainly by free AlMe3 contained in MAO. The methylzirconium cations are then formed by the reaction of rac‐2 and MAO. Small amount of MAO ([Al]/[Zr] = 40) is enough to completely activate rac‐1 to afford methylzirconium cations that can polymerize propylene. In the lab‐scale polymerizations carried out at 30°C in toluene, the rate of polymerization (Rp) shows maximum at [Al]/[Zr] = 6,250. The Rp increases as the polymerization temperature (Tp) increases in the range of Tp between 10 and 70°C and as the catalyst concentration increases in the range between 21.9 and 109.6 μM. The activation energies evaluated by simple kinetic scheme are 4.7 kcal/mol during the acceleration period of polymerization and 12.2 kcal/mol for an overall reaction. The introduction of additional free AlMe3 before activating rac‐1 with MAO during polymerization deeply influences the polymerization behavior. The iPPs obtained at various conditions are characterized by high melting point (approximately 155°C), high stereoregularity (almost 100% [mmmm] pentad), low molecular weight (MW), and narrow molecular weight distribution (below 2.0). The fractionation results by various solvents show that iPPs produced at Tp below 30°C are compositionally homogeneous, but those obtained at Tp above 40°C are separated into many fractions. © 1999 John Wiley & Sons, Inc. J Polym Sci A: Polym Chem 37: 737–750, 1999  相似文献   

7.
The influence of anhydrous ferric chloride on the catalytic properties of chloroaluminate ionic liquids catalyst for Friedel–Crafts alkylation was investigated. The catalysts were characterized by Fourier‐transform infrared (FT‐IR) (acetonitrile molecule as probe), specific gravity, and 27Al NMR. Besides, the effect of the mass ratio of FeCl3 to AlCl3, catalysts dosage, toluene/olefin molar ratio, reaction temperature, and reaction time on long‐chain alkenes alkylation were investigated thoroughly. And bromine value and high‐performance liquid chromatography (HPLC) were employed as the evaluation method for alkylation products. It was observed that the addition of anhydrous ferric chloride results in improvement in terms of Lewis acid and its catalytic recyclability. Among these catalysts studied, the catalyst modified with 1.0 wt.% anhydrous FeCl3 showed the best catalytic performance in terms of yield and stability, which can be attributed to the formation of new stronger acidic ions [Al2FeCll0]? when the added ferric chloride reacts with acidic ions [Al2Cl7]?.  相似文献   

8.
To optimize yields, the study of reaction kinetics related to the synthesis of 2‐hydroxyethylhydrazine (HEH) obtained from the alkylation of N2H4 by 2‐chloroethanol (CletOH) was carried out with and without sodium hydroxide. In both cases, the main reaction of HEH formation was followed by a consecutive, parallel reaction of HEH alkylation (or dialkylation of N2H4), leading to the formation of two isomers: 1,1‐di(hydroxyethyl)hydrazine and 1,2‐di(hydroxyethyl)hydrazine. In this study, hydrazine and hydroxyalkylhydrazine alkylations followed SN2 reactions triggered directly by CletOH or indirectly in the presence of a strong base by ethylene oxide, an intermediate compound. The kinetics was studied in diluted mediums by quantifying HEH and CletOH by gas chromatography and gas chromatography coupled with mass spectrometry (GC‐MS). The activation parameters of each reaction and the influence of a strong base present in the medium on the reaction mechanisms were established. A global mathematical treatment was applied for each alternative. It allowed modeling the reactions as a function of reagent concentrations and temperature. In the case of direct alkylation by CletOH, simulation was established for semi‐batch and batch syntheses and was confirmed in experiments for concentrated mediums (1.0 M ≤ [CletOH]0 ≤ 3.2 M and 15.7 M ≤ [N2H4]0 ≤ 18.8 M). Simulation therefore permits the prediction of the instantaneous concentration of reagents and products, in particular ethylene oxide concentration in the case of indirect alkylation, which must be as weak as possible. © 2009 Wiley Periodicals, Inc. Int J Chem Kinet 41: 382–393, 2009  相似文献   

9.
3(2‐pyridinylmethylene)‐5‐aryl‐2(3H)‐furanones and 3(3‐pyridinylmethylene)‐5‐aryl‐2(3H)‐furanones were prepared as a mixture of (E) and (Z) stereoisomers by condensing pyridine‐2‐carboxaldehyde and pyridine‐3‐carboxaldehyde with 3‐aroylpropionic acids. The reaction of the furanones 6 and 7 with anhydrous aluminium chloride in benzene led to the formation of 4,4‐diaryl‐1‐(2‐pyridinyl)but‐1,3‐diene ( 8 ) and 4,4‐diaryl‐1‐(3‐pyridinyl)but‐1,3‐diene ( 9 ) as mixtures of geometrical (E,E‐ and E,Z‐) stereoisomers via an intermolecular alkylation mode. When the reaction was carried out in tetrachloroethane as a solvent, the reaction of 6 gave 5‐arylquinoline‐7‐carboxylic acid via intramolecular alkylation mode. This may be considered as a novel method for the synthesis of quinoline derivatives. J. Heterocyclic Chem., (2011).  相似文献   

10.
Flavone ( 1 ) was easily reduced by using the electrochemical method to give two hydrodimers of 2,2′‐biflavanone(racemate) ( 5a ) and 2,2′‐biflavanone(meso) ( 5b ) and one reductive product of flavanone ( 6 ). Their yields were dependent on the nature of electrodes, the kinds of supporting electrolytes and the reaction temperature. They were found to afford higher yields of 2,2′‐biflavanone(racemate) ( 5a ) and 2,2′‐biflavanone (meso) ( 5b ) (32.4% and 24.8%, 35.8% and 13.4%, respectively,) in the reaction conditions of Pb(‐)/C(+)‐H2SO4‐7F/mol and C(‐)/C(+)‐H2SO4‐5F/mol.  相似文献   

11.
The title compounds were prepared by aldol reaction of anisaldehyde and the respective N,N‐dibenzyl glycinates. Deprotection of the nitrogen atom with Pearlman’s catalyst delivered the unprotected β‐hydroxytyrosine esters, which were further N‐protected as N,N‐phthaloyl (Phth) and N‐fluorenylmethylcarbonyloxy (Fmoc) derivatives. The Friedel–Crafts reaction with various arenes was studied employing these alcohols as electrophiles. It turned out that the facial diastereoselectivitiy depends on the nitrogen protecting group and on the ester group. The unprotected substrates (NH2) gave preferentially syn‐products but the anti‐selectivity increased when going from NHFmoc over NPhth to NBn2. If the ester substituent was varied the syn‐preference increased in the order Me <Et <iPr. The reactions were shown to be fully stereoconvergent and proceeded under kinetic product control. A model is suggested to explain the facial diastereoselectivity based on a conformationally locked benzylic cation intermediate. The reactions are preparatively useful for the N‐unprotected isopropyl ester, which gave Friedel–Crafts alkylation products with good syn‐selectivity (anti/syn=21:79 to 7:93), and for the N,N‐dibenzyl‐protected methyl ester, which led preferentially to anti‐products (anti/syn=80:20 to >95:5). Upon acetylation of the latter compound to the respective acetate, Bi(OTf)3‐catalyzed alkylation reactions became possible, in which silyl enol ethers served as nucleophiles. The respective alkylation products were obtained in high yield and with excellent anti‐selectivitiy (anti/syn≥95:5).  相似文献   

12.
Treatment of 1‐aryl‐1‐allen‐6‐enes with [PPh3AuCl]/AgSbF6 (5 mol %) in CH2Cl2 at 25 °C led to intramolecular [3+2] cycloadditions, giving cis‐fused dihydrobenzo[a]fluorene products efficiently and selectively. The reactions proceeded with initial formation of trans/cis mixtures of 2‐alkyl‐1‐isopropyl‐2‐phenyl‐1,2‐dihydronaphthalene cations B, which were convertible into the desired cis‐fused cycloadducts through the combined action of a gold catalyst and a Brønsted acid. Theoretic calculation supports the participation of the trans‐B cation as reaction intermediate. Although HOTf showed similar activity towards several 1‐aryl‐1‐allen‐6‐enes, it lacks generality for this cycloaddition reaction.  相似文献   

13.
The enantioselective ketimine–ene reaction is one of the most challenging stereocontrolled reaction types in organic synthesis. In this work, catalytic enantioselective ketimine–ene reactions of 2‐aryl‐3H‐indol‐3‐ones with α‐methylstyrenes were achieved by utilizing a B(C6F5)3/chiral phosphoric acid (CPA) catalyst. These ketimine–ene reactions proceed well with low catalyst loading (B(C6F5)3/CPA=2 mol %/2 mol %) under mild conditions, providing rapid and facile access to a series of functionalized 2‐allyl‐indolin‐3‐ones with very good reactivity (up to 99 % yield) and excellent enantioselectivity (up to 99 % ee). Theoretical calculations reveal that enhancement of the acidity of the chiral phosphoric acid by B(C6F5)3 significantly reduces the activation free energy barrier. Furthermore, collective favorable hydrogen‐bonding interactions, especially the enhanced N?H???O hydrogen‐bonding interaction, differentiates the free energy of the transition states of CPA and B(C6F5)3/CPA, thereby inducing the improvement of stereoselectivity.  相似文献   

14.
A simple five‐step synthesis of fully substituted (4RS,5RS)‐4‐aminopyrazolidin‐3‐ones as analogs of D ‐cycloserine was developed. It comprises a two‐step preparation of 5‐substituted (4RS,5RS)‐4‐(benzyloxycarbonylamino)pyrazolidin‐3‐ones, reductive alkylation at N(1), alkylation of the amidic N(2) with alkyl halides, and simultaneous hydrogenolytic deprotection/reductive alkylation of the primary NH2 group. The synthesis enables an easy stepwise functionalization of the pyrazolidin‐3‐one core with only two types of common reagents, aldehydes (or ketones) and alkyl halides. The structures of products were elucidated by NMR spectroscopy and X‐ray diffraction.  相似文献   

15.
Ring‐opening polymerization of ?‐caprolactone was carried out smoothly and effectively with constant microwave powers of 170, 340, 510, and 680 W, respectively, with a microwave oven at a frequency of 2.45 GHz. The temperature of the polymerization ranged from 80 to 210 °C. Poly(?‐caprolactone) (PCL) with a weight‐average molar mass (Mw) of 124,000 g/mol and yield of 90% was obtained at 680 W for 30 min using 0.1% (mol/mol) stannous octanoate as a catalyst. When the polymerization was catalyzed by 1% (w/w) zinc powder, the Mw of PCL was 92,300 g/mol after the reaction mixture was irradiated at 680 W for 270 min. © 2002 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 40: 1749–1755, 2002  相似文献   

16.
A cascade Michael‐alkylation reaction of diethyl benzylidenemalonates with chloroacetophenones was realized by using K2CO3 as the promoter. With this method, a series of diethyl trans‐2,3‐disubstituted 1,1‐cyclopropane‐ dicarboxylates have been facilely synthesized in moderate yields under mild conditions. The relative configurations of two typical products were confirmed by X‐ray crystallographic analysis.  相似文献   

17.
Solvothermal reactions of HgI2, 4,4′‐vinylenedipyridine, and HI in alcoholic solution (methanol, ethanol, or pentanol) gave rise to a family of organic‐inorganic hybrid complexes, formulated as [C14H16N2][I4]2– ( 1 ), [C16H20N2][HgI4] ( 2 ), and [C22H32N2][HgI4]4 ( 3 ). Single‐crystal X‐ray diffraction reveals that all three compounds are discrete structures, including the inorganic anion [I4]2– or [HgI4]2– and an organic cation, where the resulting organic cations were generated in situ alkylation reactions of 4,4′‐vinylenedipyridine with alcohols, with cleavage of the alcoholic C–O bond followed by a one‐step in situ N‐alkylation reaction of 4,4′‐vinylenedipyridine in acidic HI solution. X‐ray powder diffraction (XRD), 1H NMR and 13C NMR, energy‐dispersive X‐ray (EDS), IR, as well as UV/Vis/NIR spectroscopy, elemental analysis, and thermogravimetric analysis (TGA) were used to characterize the complexes.  相似文献   

18.
o‐Lithio yV‐methyl benzamides ( 1a‐f ) upon alkylation with ethyl methyl ketone gave (±)‐3‐ethyl‐3‐methyl phthalides ( 2a‐f ), which upon treatment with concentrated H2SO4 or anhydrous A1C13 furnished corresponding 3,3‐dimethyl‐3,4‐dihydroisocoumarins ( 3a‐f ) and 3‐methyl mellein ( 3g ).  相似文献   

19.
The first Lewis acid catalyzed asymmetric Friedel–Crafts alkylation reaction of ortho‐hydroxybenzyl alcohols with C3‐substituted indoles is described. A chiral N,N′‐dioxide Sc(OTf)3 complex served not only to promote formation of ortho‐quinone methides (o‐QMs) in situ but also induced the asymmetry of the reaction. This methodology enables a novel activation of ortho‐hydroxybenzyl alcohols, thus affording the desired chiral diarylindol‐2‐ylmethanes in up to 99 % yield and 99 % ee. A range of functional groups were also tolerated under the mild reaction conditions. Moreover, this strategy gives concise access to enantioenriched indole‐fused benzoxocines.  相似文献   

20.
Acid‐catalyzed Friedel–Crafts alkylation of 1,3‐dicarbonyl compounds with electrophilic alcohols, is known to be an effective C? C bond forming reaction. However, until now, this reaction has not been amenable for α‐alkylation of aryl methyl ketones because of the notoriously low nucleophilicities of these compounds. Therefore, α‐alkylation of aryl methyl ketone relies on precious metal catalysts and also, the use of primary alcohols is mandatory. In this study, we found that a system composed of a Fe(OTf)3 catalyst and chlorobenzene solvent is sufficient to promote the title Friedel–Crafts reaction by using benzhydrols as electrophiles. 3,4‐Dihydro‐9‐(2‐hydroxy‐4,4‐dimethyl‐6‐oxo‐1‐cyclohexen‐1‐yl)‐3,3‐dimethyl‐xanthen‐1(2 H)‐one was also applicable as an electrophile in this type of benzylation reaction. On the basis of this result, a three‐component reaction of salicylaldehyde, dimedone, and aryl methyl ketone was also developed, and this provided an efficient way for the synthesis of densely substituted 4H‐chromene derivatives.  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