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 31 毫秒
1.
Friedel–Crasfts alkylation reactions of α,β-unsaturated butyric aldehydes with N,N-dimethyl-3-anisidine catalyzed by a (2S,5S)-5-benzyl-2-tert-butyl-3-methylimidazolidin-4-one HCl salt have been carried out at the PCM(CH2Cl2)/B3LYP/6-311++G(d,p)//B3LYP/6-31G(d) level. Three reaction processes have been characterized: (I) the formation of an iminium ion intermediate; (II) the 1,4-iminium addition of the iminium ion; and (III) the hydrolysis of the addition product. Moreover, Path 1-1 is the favorable channel in the formation of the iminium ion. From the point of view of energy, the enantioselectivity is controlled by the carbon–carbon bond formation step that is involved in both the intermediate M4 and the transition state TS4. The highest energy barrier of the reaction is the H2 proton transfer from the O10 atom of a water molecule to the N1 atom of the catalyst in the hydrolysis process, which is 23.4 kcal/mol. The presented calculated results may be helpful in understanding the experimental product distribution for the title reaction, and provide a general model to help explain the mechanisms of similar reactions.  相似文献   

2.
《Tetrahedron: Asymmetry》2001,12(17):2389-2393
Oxidation of various 1-thioglycopyrano- and furanosides with H2O2/Ac2O/SiO2 was performed in CH2Cl2/aprotic perfluorinated solvent mixtures (100:0 to 1:19, v/v). Reactions appeared much faster at a solvent ratio of 1:1 without significant over-oxidation to sulfones and modification of diastereoselectivity. Nevertheless, the nature of both the glycosyl moiety and the protecting group influenced the (SS):(RS) product ratio.  相似文献   

3.
The present study describes the total synthesis of 1,2,3,4-tetrahydroquinoline alkaloids (±)-galipinine, (±)-cuspareine, (±)-galipeine and (±)-angustureine, in three steps and high yields (78%, 76%, 74%, and 66%, respectively) from common aldehyde and the ylide respectives. The key step of this approach is based on an unusual Wittig reaction by using the phase transfer medium (aq. NaOH/CH2Cl2 1:1 or t-BuOK/t-BuOH/CH2Cl2 1:1), affording olefinic intermediates in high yields.  相似文献   

4.
Hydrozirconation of terminal alkynes in 1-butyl-3-methylimidazolium hexafluorophosphate ([bmim][PF6]) at 30 °C gave highly regio- and stereoselectively (E)-vinylzirconium complexes, which underwent a cross-coupling reaction with aryl halides in the presence of Pd(PPh3)4 to afford (E)-1,2-disubstituted ethenes in good to high yields. Our system not only avoids the use of easily volatile THF or CH2Cl2 as solvent but also solves the basic problem of palladium catalyst reuse.  相似文献   

5.
The absolute configuration of the title acid 2 has been determined to be S by X-ray crystallography. Thus, decarboxylation of 2 produces (S)-(+)-halothane with 99% retention of configuration. This behavior is compared to other stereoselective decarboxylation reactions of α-haloacids from the literature that also gave high degrees of retention of configuration when in the form of their quaternary ammonium salts, which contain one proton. The proton of the ammonium salt is necessary in order to protonate the anionic intermediate formed from decarboxylation. In the absence of this relatively acidic proton, we had previously found that using triethylene glycol (TEG) as both the solvent and proton source for the decarboxylation reaction of acid 2 caused poor stereoselectivity. This was in contrast to 1,2,2,2-tetrafluoro-1-methoxypropionic acid 6, which showed a high degree of retention of configuration in TEG. In order to rationalize this differing behavior, we report DFT studies at PCM-B3LYP/6-31++G7 level of theory (the results were additionally confirmed with 6-311++G7 and aug-cc-pVDZ basis sets). The energy barrier to inversion of configuration of the anionic reaction intermediate 11 of acid 2 is 10.23 kcal/mol. However, we find that the anionic intermediate 10 from acid 6 would rather undergo β-elimination instead of inversion of configuration. Thus the planar transition state required for inversion of configuration is never reached, regardless of the rate of proton transfer to the anion.  相似文献   

6.
Tricycle 6, containing the CD ring of taxol, is constructed from (S)-(+)-carvone in 21 steps involving a Diels-Alder reaction with isoprene, a Baeyer-Villiger oxidation, an Oppenaurer oxidation and Meerwein-Ponndorf-Verley reduction, a stereospecific Grignard addition, and an intramolecular SN2 reaction as the key steps.  相似文献   

7.
Stereoselective synthesis of (3S,4S)- and (3R,4R)-series of 3,4-dihydroxyglutamic acids was investigated. The key reaction in this synthesis is asymmetric reduction of meso-imide derived from meso-tartaric acid. Lewis acid-promoted cyanation of the obtained optically active lactam via the acyliminium intermediate followed by standard deprotection procedure afforded the desired 3,4-dihydroxyglutamic acids.  相似文献   

8.
Imidazolines were prepared in one-pot operation from aldehydes and diamines through oxidation of aminal intermediates by NBS. This method could be applied to various aromatic and aliphatic aldehydes and N-nonsubstituted and N-monosubstituted 1,2-diamines. Furthermore, it was found that CH2Cl2 could be altered to TBME, a more environmentally friendly solvent, in the reaction using N-nonsubstituted 1,2-diamines. The reaction conditions were very mild and chemoselective.  相似文献   

9.
(Ra)-(R)2-2,2′-Bis(1-hydroxy-1H-perfluorooctyl)biphenyl ((Ra)-(R)2-1c), which is an axially dissymmetric ligand with two chiral centers, works as a good chiral auxiliary for asymmetric aldol reaction. Thus, the reaction of monopropanoyl ester of 1c (2) with benzaldehyde in the presence of triethylamine and titanium(IV) chloride gave (2R),(3S)- and (2R),(3R)-3-hydroxy-2-methyl-3-phenylpropanoic acid esters (3a) in an approximate ratio of 4:1 in a total high yield. This result shows that stereoselectivity at 2-position is quite high, while that at 3-position is moderate. Both isomers were easily separated by column chromatography. Methanolysis of the separated isomers gave nearly quantitative recovery of 1c by extraction with a fluorous solvent without any loss of ee, while methyl (2R),(3S)- or (2R),(3R)-3-hydroxy-2-methyl-3-phenylpropanoates were obtained by CH2Cl2 extraction quantitatively in >99% ee. Aldol reaction of 2 with various aldehydes gave similar results.  相似文献   

10.
By using of precise catalytic amount of N-methylpyrrolidine (5 mol %) and Ba(OH)2 (1.5 mol %) in H2O/CH3OH 5/1 or CH3OH/CH2Cl2 3/1 solvent mixtures at T=0 °C a Morita-Baylis-Hillman derivatives could be obtained in good to excellent yield from 2-cyclopenten-1-one, 2-cyclohexen-1-one and formaldehyde and diverse aryl aldehydes after suitable reaction time.  相似文献   

11.
Various α-keto-1,3,4-oxadiazole derivatives were synthesized through a sequential intermolecular dehydrochlorination/intramolecular aza-Wittig reaction of carboxylic acids and imidoyl chloride intermediates, which were generated by isocyanide-Nef reaction of acyl chlorides and (N-isocyanimine) triphenylphosphorane (1) in CH2Cl2 at room temperature.  相似文献   

12.
Reaction of Adipic Acid Diamide with Phosphorus Pentachloride The reaction of adipamide (I) with phosphorus pentachloride in a solvent leads to (Cl3P?NCCl2CCl2CH2)2 (II). The stages of the reaction are: 1. chlorination of the keto and methylen groups 2. formation of the ? N?PCl3 group. This result is a supplement of the existing conception about the course of the reaction of carboxylic acid amides with phosphorus pentachloride. The reaction of (I) with PCl5 without any solvent has been reproduced and the course of reaction has also been investigated. This reaction gives mainly NC(CH2)4CN. The resulting product of a careful hydrolysis of (II) is (Cl2OPN?CClCl2CH2)2. A total hydrolysis gives back (I).  相似文献   

13.
In a study using UV photoelectron spectroscopy (PES) of the atmospherically relevant reaction
CH3SCH3 + Cl2 → CH3SCH2Cl + HCl
bands associated with a reaction intermediate have been observed. These have been assigned to ionization of the covalently bound molecule (CH3)2SCl2 on the basis of the intensity of the observed bands as a function of reaction time, molecular orbital calculations of vertical ionization energies and evidence from infrared spectroscopy.  相似文献   

14.
The kinetics of oxidation of 1-octene and heptanal by 18-crown-6-ether-solubilized KMnO4 in benzene and CH2Cl2 have been investigated. In benzene, the oxidation of 1-octene is first order with respect to the oxidant and zero order with respect to the substrate, whereas in CH2Cl2 the reaction is first order with respect to both substrate and oxidant. The reaction of heptanal followed different kinetics being first order with respect to both substrate and oxidant, regardless of whether benzene or CH2Cl2 was employed as the solvent. The values of activation energy E a, standard enthalpy H *, standard entropy change S *, and standard free energy G *, for the reaction, are reported. Mechanistic pathways for the studied reactions are also proposed.  相似文献   

15.
Evidence is provided that in a gas-solid photocatalytic reaction the removal of photogenerated holes from a titania (TiO2) photocatalyst is always detrimental for photocatalytic CO2 reduction. The coupling of the reaction to a sacrificial oxidation reaction hinders or entirely prohibits the formation of CH4 as a reduction product. This agrees with earlier work in which the detrimental effect of oxygen-evolving cocatalysts was demonstrated. Photocatalytic alcohol oxidation or even overall water splitting proceeds in these reaction systems, but carbon-containing products from CO2 reduction are no longer observed. H2 addition is also detrimental, either because it scavenges holes or because it is not an efficient proton donor on TiO2. The results are discussed in light of previously suggested reaction mechanisms for photocatalytic CO2 reduction. The formation of CH4 from CO2 is likely not a linear sequence of reduction steps but includes oxidative elementary steps. Furthermore, new hypotheses on the origin of the required protons are suggested.  相似文献   

16.
《Tetrahedron: Asymmetry》2006,17(2):281-286
The stereoselective preparation of a new Meyers’ bicyclic lactam-bridged biaryl 1, highly structurally related to circumdatins, benzomalvins and asperlicins, is reported. Using the popular Meyers’ diastereoselective lactamization, under dehydrating conditions (CH2Cl2/reflux/MgSO4), trans-(aS,R,S)-1 was obtained in a rather modest yield of 25% and an excellent diastereoselectivity >95%. An alternative procedure making use of the activating agents of carboxylic acid (Mukaiyama reagent and FEP) allowed the lactamization process to take place under milder conditions (CH2Cl2/20 °C) affording trans-(aS,R,S)-1 in fairly good yields (50%–85%) and in up to 65% de.  相似文献   

17.
The reaction of S4N4Cl2 with CH3OH gives S4N4(OCH3)2, a simple dimethoxoderivative of S4N4. Its overall geometry is analogous to other compounds of the S4N4X2 type. The chlorination of S4N4(OCH3)2 leads to the oxidation of one sulfur atom to SVI and CH3OS4N4(O)Cl is formed. The compounds were characterized by ir spectroscopy and their crystal structures were determined from single crystal diffraction data collected at ?153°C. The presence of SVI in the molecule of CH3OS4N4(O)Cl is manifested by a marked shortening of the bonds formed by this atom as compared with S4N4Cl2 and S4N4(OCH3)2.  相似文献   

18.
A high oxidation state alkylnitridoosmium complex, [Os(N)(CH2SiMe3)4][NBun4] acts as a nucleophile in reactions with alkyl halides. Alkylimido complexes, Os(NR)(CH2SiMe3)4, are produced. The reaction between [Os(N)(CH2SiMe3)4] [NBun4] and MeI is second order with k2= 9.5 x 10−5 sect̄1 M−1 at 23°C in CD2Cl2 under pseudo first order conditions. The entropy of activation, ΔS, was found to be −10.6 ± 0.5 cal M−1 K−1 and the enthalpy of activation, ΔH, was found to be 19.6 ± 0.2 kcal M−1. The reaction proceeds faster in polar, non-coordinating solvents than in either non-polar solvents or in solvents which can coordinate to the osmium center.  相似文献   

19.
Liquid-crystalline solutions of cellulose triacetate (CTA) in trifluoroacetic acid (TFA)–CH2Cl2, TFA–1.2-dichloroethane (1,2-DCE) solvent mixtures were examined by means of PMR spectroscopy. CTA forms both cholesteric and nematic phases in these solvents depending on the CTA concentration. In cholesteric solutions the CH2Cl2 signal is initially a singlet and then splits into a doublet. The time dependence of the splitting and the effect of CTA concentration are reported. The results suggest that the cholesteric phase slowly changes into a nematic phase in the magnetic field. The splitting of the CH2Cl2 proton signal into a doublet and the 1,2-DCE signal into a quartet are due to direct magnetic dipole-dipole interactions. Rotation of the sample in the magnetic field results in the disappearance of the doublet or quartet and suggests that the solvent molecules are originally oriented in the direction of the magnetic field. In the biphasic region, immediate splitting of the CH2Cl2 proton signal suggests that the anisotropic phase is nematic.  相似文献   

20.
Synthesis and Crystal Structure of (PPh4)2[Mo2(S2)2Cl8] · 2 CH3CN and its Topotactic Transformation to (PPh4)2[Mo2(S2)2Cl8] MoS2Cl3 was prepared from molybdenum and S2Cl2 at 200 °C. Its reaction with PPh4Cl in acetonitrile yielded (PPh4)2[Mo2(S2)2Cl8] · 2 CH3CN. In vacuum or upon warming, it loses the acetronitrile without degradation of the crystals. According to the X-ray crystal structure determinations both compounds, with and without acetonitrile, are triclinic. They contain the same [Cl4Mo(μ-S2)2MoCl4]2– ions, in which the Mo atoms are joined by two disulfido groups and an Mo–Mo bond. Details of the crystal packings and their topotactic transformation are given.  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