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.
The treatment of 3,3a,4,5‐tetrahydro‐3‐aryl‐2‐phenyl‐2H‐benzo[g]indazoles 4 with I2/DMSO led to the oxidation of the five‐member rings ( 5 ) as well as the iodination of N‐phenyl moieties along with oxidation of the five‐member rings ( 6 ). However, the reactions of 4 with CuCl2/DMSO gave only compounds 5 . The reaction of N‐bromosuccinimide (NBS) with compounds 4 resulted in fully aromatization along with bromination at C‐5 of the indazole rings ( 7 ). The indazole six‐member rings in compounds 5 and 6 also underwent aromatization along with bromination by using NBS ( 7 and 8 ).  相似文献   

2.
The kinetics of Ru(III)‐catalyzed and Hg(II)‐co‐catalyzed oxidation of D‐glucose (Glc) and cellobiose (Cel) by N‐bromoacetamide (NBA) in the presence of perchloric acid at 40 °C have been investigated. The reactions exhibit the first order kinetics with respect to NBA, but tend towards the zeroth order to higher NBA. The reactions are the first order with respect to Ru(III) and are fractional positive order with respect to [reducing sugar]. Positive effect of Cl? and Hg(OAc)2 on the rate of reaction is also evident in the oxidation of both reducing sugars. A negative effect of variation of H+ and acetamide was observed whereas the ionic strength (µ) of the medium had no influence on the oxidation rate. The rate of reaction decreased with the increase in dielectric constant and this enabled the computation of dAB, the size of the activated complex. Various activation parameters have been evaluated and suitable explanation for the formation of the most reactive activated complex has been given. The main products of the oxidation are the corresponding arabinonic acid and formic acid. HOBr and [RuCl3(H2O)2OH]? were postulated as the reactive species of oxidant and catalyst respectively. A common mechanism, consistent with the kinetic data and supported by the observed effect of ionic strength, dielectric constant and multiple regression analysis, has been proposed. Formation of complex species such as [RuCl3·S·(H2O)OH]? and RuCl3·S·OHgBr·OH during the course of reaction was fully supported by kinetic and spectral evidences.  相似文献   

3.
A reaction of diethyl 2,4,6‐trimethyl‐1,4‐dihydropyridine‐3,5‐dicarboxylate with 1, 2, and more equivalents of N‐bromosuccinimide (NBS) in methanol was investigated by NMR spectroscopy at a temperature interval ranging from 25 to 40°C. The reaction was found to proceed through several steps. The structures of the intermediates diethyl 3‐bromo‐2,4,6‐trimethyl‐3,4‐dihydropyridine‐3,5‐dicarboxylate, diethyl 3‐bromo‐2‐methoxy‐2,4,6‐trimethyl‐1,2,3,4‐tetrahydropyridine‐3,5‐dicarboxylate, and diethyl 3,5‐dibromo‐2‐methoxy‐2,4,6‐trimethyl‐2,3,4,5‐tetrahydropyridine‐3,5‐dicarboxylate were identified by multinuclear (1H, 13C, and 15N) NMR spectral data. The optimal structures of all species participating in the reaction as well as changes in their relative energies along with the proposed pathway of the reaction were analyzed by quantum‐chemical calculations. The mechanism of bromination of diethyl 2,4,6‐trimethyl‐1,4‐dihydropyridine‐3,5‐dicarboxylate with NBS in methanol was found to favor the bromination in the 2,6‐methyl side chains as the only products in full agreement with experimental observations.  相似文献   

4.
Selective bromination of γ‐methylene C(sp3)−H bonds of aliphatic amides and δ‐methylene C(sp3)−H bonds of nosyl‐protected alkyl amines are developed using NBS as the brominating reagent and catalytic amount of CuII/phenanthroline complexes as the catalyst. Aryl and benzylic C−H bonds at other locations remain intact during this directed radical abstraction reaction.  相似文献   

5.
β‐Carbolines ( 1–5 ) undergo electrophilic aromatic substitution with N‐bromosuccinimide under different experimental conditions. Although 6‐bromo‐nor‐harmane ( la ) obtained by bromination of nor‐harmane ( 1 ) was isolated and fully characterized sometime ago, the other bromoderivatives of nor‐harmane ( 1b‐1e ) and harmane ( 2a‐2e ) were partially described as part of the reaction mixtures. The preparation and subsequent isolation, purification and full characterization of 1b, 1c, 1d, 1e, 2a, 2b, 2c, 2d, 2e are reported (mp, R f, 1H‐nmr, 13C‐nmr and ms) together with the preparation, isolation and charaterization, for the first time, of the bromoderivatives obtained from harmine ( 3a‐3e ), harmol ( 4a, 4b ) and 7‐acetylharmol ( 5a‐5c ). As brominating reagent N‐bromosuccinimide and N‐bromosuccinimide‐silica gel in dichloromethane and in chloroform as well as the β‐carboline ‐ N‐bomosuccinimide solid mixture have been used and their uses have been compared. Semiempirical AMI and PM3 calculations have been performed in order to predict reactivity in terms of the energies of HOMO, HOMO‐LUMO difference and in terms of the charge density of β‐carbolines ( 1–5 ) and bromo‐β‐carbolines ( 1a‐1e, 2a‐2e, 3a‐3e, 4a, 4b, 5a, 5b and 5c ) (Scheme 1). Theoretical and experimental results are discussed briefly.  相似文献   

6.
The preparation, isolation and characterization by elemental analysis and 1H‐NMR, 13C‐NMR, and MS data of the bromo derivatives of N‐substituted carbazoles, i.e., of 9‐methyl‐9H‐carbazole ( 1 ), 9‐phenyl‐9H‐carbazole ( 2 ), 9‐benzyl‐9H‐carbazole ( 3 ), 2‐methoxy‐9‐methyl‐9H‐carbazole ( 4 ), and of C‐substituted carbazoles, i.e., of 2‐(acetyloxy)‐9H‐carbazole ( 5 ) and 3‐nitro‐9H‐carbazole ( 6 ), are reported, in part for the first time. As brominating reagents, N‐bromosuccinimide (NBS) or NBS/silica gel in CH2Cl2, NBS in AcOH, KBrO3/KBr in EtOH doped with a catalytic amount of H2SO4, or KBrO3/KBr in AcOH were employed, and their uses were compared. Semi‐empirical PM3 calculations were performed to predict the reactivity of the N‐substituted and C‐substituted carbazoles and of their bromo derivatives and found to verify the experimental results. The UV‐absorption and fluorescence and phosphorescence emission spectra of the bromocarbazole derivatives in MeCN solution at 298 K and in a solid matrix at 77 K were compared with those of the corresponding carbazoles 1 – 6 . The dynamic properties of the lowest excited singlet and triplet states (τf, τp, ?f, and ?p) were measured under the same experimental conditions. The intramolecular spin–orbital‐coupling effect of the Br‐atom and NO2 group on the spectroscopic data, photophysical parameters, and on the photo reactivity were also briefly analyzed.  相似文献   

7.
An efficient bromination protocol for the synthesis of α-bromo-β-keto esters has been developed. In PEG-400 (poly(ethylene glycol-400)), a variety of β-keto esters were treated with NBS (N-bromosuccinimide) at room temperature to selectively afford the corresponding α-monobromination products in excellent yields. It is noteworthy that the reaction was conducted under mild, environmentally benign and catalyst-free conditions.  相似文献   

8.
The oxidation of 4‐methyl‐3‐thiosemicarbazide (MTSC) by bromate and bromine was studied in acidic medium. The stoichiometry of the reaction is extremely complex, and is dependent on the ratio of the initial concentrations of the oxidant to reductant. In excess MTSC and after prolonged standing, the stoichiometry was determined to be H3CN(H)CSN(H)NH2 + 3BrO3? → 2CO2 + NH4+ + SO42? + N2 + 3Br? + H+ (A). An interim stoichiometry is also obtained in which one of the CO2 molecules is replaced by HCOOH with an overall stoichiometry of 3H3CN(H)CSN(H)NH2 + 8BrO3? → CO2 + NH4+ + SO42? + HCOOH + N2 + 3Br? + 3H+ (B). Stoichiometry A and B are not very different, and so mixtures of the two were obtained. Compared to other oxidations of thiourea‐based compounds, this reaction is moderately fast and is first order in both bromate and substrate. It is autocatalytic in HOBr. The reaction is characterized by an autocatalytic sigmoidal decay in the consumption of MTSC, while in excess bromate conditions the reaction shows an induction period before autocatalytic formation of bromine. In both cases, oxybromine chemistry, which involves the initial formation of the reactive species HOBr and Br2, is dominant. The reactions of MTSC with both HOBr and Br2 are fast, and so the overall rate of oxidation is dependent upon the rates of formation of these reactive species from bromate. Our proposed mechanism involves the initial cleavage of the C? N bond on the azo‐side of the molecule to release nitrogen and an activated sulfur species that quickly and rapidly rearranges to give a series of thiourea acids. These thiourea acids are then oxidized to the sulfonic acid before cleavage of the C? S bond to give SO42?, CO2, and NH4+. © 2002 Wiley Periodicals, Inc. Int J Chem Kinet 34: 237–247, 2002  相似文献   

9.
A H2O2-HBr system and N-bromosuccinimide in an aqueous medium were used as a ‘green’ approach to electrophilic and radical bromination. Several activated and less activated aromatic molecules, phenylsubstituted ketones and styrene were efficiently brominated ‘on water’ using both systems at ambient temperature and without an added metal or acid catalyst, whereas various non-activated toluenes were functionalized at the benzyl position in the presence of visible light as a radical activator. A comparison of reactivity and selectivity of both brominating systems reveals the H2O2-HBr system to be more reactive than NBS for benzyl bromination and for the bromination of ketones, while for electrophilic aromatic substitution of methoxy-substituted tetralone it was higher for NBS. Also, higher yields of brominated aromatics were observed when using H2O2-HBr ‘on water’. Bromination of styrene reveals that not just the structure of the brominating reagent but the reaction conditions: amount of water, organic solvent, stirring rate and interface structure, play a key role in defining the outcome of bromination (dibromination vs bromohydroxylation). In addition, mild reaction conditions, a straightforward isolation procedure, inexpensive reagents and a lower environment impact make aqueous brominating methods a possible alternative to other reported brominating protocols.  相似文献   

10.
Kinetic investigations on Ru(III)‐catalyzed oxidation of cyclopentanol and cyclohexanol by acidic solution of N‐bromoacetamide (NBA) in the presence of mercury(II) acetate as a scavenger have been carried out in the temperature range of 30–45°C. Similar kinetics was followed by both the cyclic alcohols. First‐order kinetics in the lower concentration range of NBA was observed to tend to zero order at its higher concentrations. The reaction exhibits a zero‐order rate dependence with respect to each cyclic alcohol, while it is first order in RuIII. Increase in [H+] and [Cl?] showed positive effect, while successive addition of acetamide exhibited negative effect on the reaction rate. Insignificant effect of sodium perchlorate, D2O, and mercury(II) acetate on the reaction velocity was observed. Cationic bromine has been proposed as the real oxidizing species. Various thermodynamic parameters have been computed. A suitable mechanism in agreement with the kinetic observations has been proposed. © 2005 Wiley Periodicals, Inc. Int J Chem Kinet 37: 275–281, 2005  相似文献   

11.
N‐Allyl, N‐cinnamyl, and N‐(3‐trimethylsilyl)propargyl derivatives of 4‐isopropyl‐5,5‐diphenyloxazolidin‐2‐one (DIOZ) are prepared by lithiation of the parent DIOZ (with BuLi in THF) and reaction with the corresponding bromides (Scheme 1). Lithiation in the same solvent, with deprotonation by BuLi on the allylic or propargylic CH2 group at dry‐ice temperature, provides colorful solutions, which are either combined with aldehydes or ketones directly or after addition (with or without warming) of (Me2N)3TiCl or (i‐PrO)3TiCl. Conditions have thus been elaborated under which all three types of conjugated lithium compounds react in the γ‐position with respect to the oxazolidinone N‐atom: carbamoyl derivatives of enamines and allenyl amines are formed in yields ranging from 60 to 80% and with diastereoselectivities up to 98% (Schemes 2–5). The C=C bond of the N‐hydroxyalkenyl groups has (Z)‐configuration (products 5 and 8 ), the allene chirality axis has (M)‐configuration (products 9 ), and the addition to aldehydes and unsymmetrical ketones has taken place preferentially from the Si face. A mechanistic model is proposed that is compatible with the stereochemical outcome (assuming kinetic control and disregarding the presence of Li and Ti species in the reaction mixture; cf. L, M in Fig. 4). Hydrolysis of the enamine derivatives leads to lactols, oxidizable to γ‐lactones, with recovery of the crystalline oxazolidinone, as demonstrated in three cases (Scheme 6). Thus, the application of chiral oxazolidinone auxiliaries (cf. Figs. 1 and 2) has been extended to the overall enantioselective preparation of homoaldols.  相似文献   

12.
Systems based on organic complexes of tribromide anion generate upon dissolution in water nucleophile–oxidant couple HOBr/BrO and accelerate hydrolysis of ethyl 4-nitrophenyl ethylphosphonate, diethyl 4-nitrophenyl phosphate, and 4-nitrophenyl p-toluenesulfonate by a factor of 15–90 in the presence of cationic surfactant micelles. As in water, hypobromite ion in surfactant micelles acts as α-nucleophile, and the magnitude of the α-effect almost does not change in going from water to micelles. Micellar effects of surfactants are determined by the nucleophilicity of hypobromite ion in surfactant micelles and by solubilization of the substrate and BrO, which largely depend on the counterion concentration in the micelle surface layer. The main factor responsible for the observed acceleration is increased reactant concentration in the micellar pseudophase.  相似文献   

13.
Treatment of 2‐(1‐aryl‐3‐propynoyl) anisoles 1 with N‐chlorosuccinimide (NCS) or N‐bromosuccinimide (NBS) gave the 3‐halogenated flavones and their related molecules in moderate yields.  相似文献   

14.
Maleic acid (MA) in aqueous sulfuric acid undergoes catalytic isomerization in the presence of small amounts of Cerium(IV) ion and N-bromosuccinimide (NBS) or N-bromoacetamide (NBA). The rate of isomerization is very fast even at room temperature and the yield is quite acceptable. The rate of isomerization depends on the relative amounts of MA, Ce(IV), NBS, NBA, and H2SO4. However, maleic acid has greater effect on the final yield. Sulfuric acid exhibits more chemical effect than physical effects. The competitive redox reactions of Ce(IV), NBS, and NBA with MA limit the yield of isomerization to about 85%. In the vicinity of room temperature, a raise of five degrees in temperature nearly doubles the rate of isomerization. Acrylamide shows inhibitive effect on the isomerization. The rate of hydrolysis of NBS or NBA in aqueous acidic solution depends on the concentrations of hydrogen ion, and NBS or NBA itself. The rate of hydrolysis of NBA is much faster than that of NBS. Mechanism involving bromine atom as catalyst is proposed to explain experimental results.  相似文献   

15.
The crystal structures of the p‐bromo­benzoates of cis‐4‐oxa‐1‐decalinyl (C16H19BrO3), trans‐4‐oxa‐1‐decalinyl (C16H19­BrO3), N‐benzyl‐cis‐4‐aza‐1‐decalinyl (C23H26BrNO2), N‐benzyl‐trans‐4‐aza‐1‐decalinyl (C23H26BrNO2) and trans‐4‐thia‐1‐decalinyl (C16H19BrO2S) (decalin is per­hydro­naphthalene) have been determined as part of a study directed at predicting and interpreting the π‐facial selectivities of diastereotopic ketones in reactions with nucleophiles. All five structures are composed of mol­ecules that are separated by normal van der Waals distances. In all five structures, the heterocyclic and cyclo­hexyl rings adopt chair conformations, and the p‐bromo­benzoate groups are planar.  相似文献   

16.
With regard to crystal engineering, building block or modular assembly methodologies have shown great success in the design and construction of metal–organic coordination polymers. The critical factor for the construction of coordination polymers is the rational choice of the organic building blocks and the metal centre. The reaction of Zn(OAc)2·2H2O (OAc is acetate) with 3‐nitrobenzoic acid (HNBA) and 4,4′‐bipyridine (4,4′‐bipy) under hydrothermal conditions produced a two‐dimensional zinc(II) supramolecular architecture, catena‐poly[[bis(3‐nitrobenzoato‐κ2O,O′)zinc(II)]‐μ‐4,4′‐bipyridine‐κ2N:N′], [Zn(C7H4NO4)2(C10H8N2)]n or [Zn(NBA)2(4,4′‐bipy)]n, which was characterized by elemental analysis, IR spectroscopy, thermogravimetric analysis and single‐crystal X‐ray diffraction analysis. The ZnII ions are connected by the 4,4′‐bipy ligands to form a one‐dimensional zigzag chain and the chains are decorated with anionic NBA ligands which interact further through aromatic π–π stacking interactions, expanding the structure into a threefold interpenetrated two‐dimensional supramolecular architecture. The solid‐state fluorescence analysis indicates a slight blue shift compared with pure 4,4′‐bipyridine and HNBA.  相似文献   

17.
The Ti(TADDOLato) complexes dichloro[(4R,5R)‐2,2‐dimethyl‐α,α,α′,α′‐tetraphenyl‐1,3‐dioxolane‐4,5‐dimethanolato(2−)‐O,O′]titanium ((R)‐ 1a ) and dichloro[(4R,5R)‐2,2‐dimethyl‐α,α,α′,α′‐tetra(naphthalen‐1‐yl)‐1,3‐dioxolane‐4,5‐dimethanolato(2−)‐O,O′]titanium ((R)‐ 1b ) are efficient catalysts for the electrophilic enantioselective chlorination and bromination of β‐keto esters with N‐chlorosuccinimide (NCS) and N‐bromosuccinimide (NBS), respectively. With 5 mol‐% of catalyst at room temperature an enantioselectivity of up to 88% ee could be obtained for the chlorination reaction. Under comparable conditions, bromination reactions are slower and less stereoselective.  相似文献   

18.
Amino acid‐derived novel norbornene derivatives, N,N′‐(endo‐bicyclo[2.2.1] hept‐5‐en‐2,3‐diyldicarbonyl) bis‐L ‐alanine methyl ester (NBA), N,N′‐(endo‐bicyclo[2.2.1]hept‐5‐en‐2,3‐diyldicarbonyl) bis‐L ‐leucine methyl ester (NBL), N,N′‐(endo‐bicyclo[2.2.1]hept‐5‐en‐2,3‐diyldicarbonyl) bis‐L ‐phenylalanine methyl ester (NBF) were synthesized and polymerized using the Grubbs 2nd generation ruthenium (Ru) catalyst. Although NBA, NBL, and NBF did not undergo homopolymerization, they underwent copolymerization with norbornene (NB) to give the copolymers with Mn ranging from 5200 to 38,100. The maximum incorporation ratio of the amino acid‐based unit was 9%, and the cis contents of the main chain were 54–66%. © 2006 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 44: 5337–5343, 2006  相似文献   

19.
Polymers containing o‐nitrobenzyl esters are promising for preparation of light sensitive materials. o‐Nitrobenzyl methacrylate has already been polymerized by controlled ATRP or RAFT. Unfortunately, the radical polymerization of o‐nitrobenzyl acrylate (NBA) was not controlled until now due to inhibition and retardation effects coming from the nitro‐aromatic groups. Recent developments in the Single Electron Transfer–Living Radical Polymerization (SET–LRP) provide us an access to control this NBA polymerization and living character of this NBA SET–LRP is demonstrated. Effects of CuBr2 and ligand concentrations, as well as Cu(0) wire length on SET–LRP kinetics are shown presently. A first‐order kinetics with respect to the NBA concentration is observed after one induction period. SET–LRP proceeds with a linear evolution of molecular weight and a narrow distribution. High initiation efficiency close to 1 and high chain‐end functionality (~93%) are reached. Chain extension of poly(o‐nitrobenzyl acrylate) is realized with methyl acrylate (MA) to obtain well defined poly(o‐nitrobenzyl acrylate)‐b‐poly(methyl acrylate) (PNBA‐b‐PMA). Finally, light‐sensitive properties of PNBA are checked upon UV irradiation. © 2014 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2014 , 52, 2192–2201  相似文献   

20.
The bromination of dimethyl 8‐methoxy‐1,6,10‐trimethylheptalene‐4,5‐dicarboxylate ( 6 ; Scheme 2) with N‐bromosuccinimide (NBS) in N,N‐dimethylformamide (DMF) leads in acceptable yields to the corresponding 9‐bromoheptalenedicarboxylate 10 (Table 1). Ether cleavage of 6 with chlorotrimethylsilane (Me3SiCl)/NaI results in the formation of oxoheptalenedicarboxylate 13 in good yield (Scheme 4). The latter can be acetyloxylated to the (acetyloxy)oxoheptalenedicarboxylate 14 with Pb(OAc)4 in benzene (Scheme 5). Oxo derivative 14 , in turn, can be selectively O‐methylated with dimethyl sulfate (DMS) in acetone to the (acetyloxy)methoxyheptalenedicarboxylates 15 and 15′ (Scheme 6). The AcO group of the latter can be transformed into a benzyl or methyl ether group by treatment with MeONa in DMF, followed by the addition of benzyl bromide or methyl iodide (cf. Scheme 9). Reduction of the ester groups of dimethyl 7,8‐dimethoxy‐5,6,10‐trimethylheptalene‐1,2‐dicarboxylate ( 25′ ) with diisobutylaluminium hydride (DIBAH) in tetrahydrofuran (THF) leads to the formation of the corresponding dimethanol 26′ , which can be cyclized oxidatively (IBX, dimethyl sulfoxide) to 8,9‐dimethoxy‐6,7,11‐trimethylheptaleno[1,2‐c]furan ( 27 ; Scheme 11).  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