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 0 毫秒
1.
Solvent-free ester aminolysis was studied under microwave or conventional heating either in the absence of base or induced by KOtBu with or without a phase transfer agent. The specific microwave effects were shown to be dependent on the conditions and discussed in terms of relative polarities of ground and transition states.  相似文献   

2.
3.
《Tetrahedron: Asymmetry》2006,17(8):1264-1274
In a previous work, several 2-phenylcycloalkanamines were subjected to aminolysis catalysed by the lipase B from Candida antarctica (CAL-B). In these processes, the size of the cycle and the stereochemistry of the stereogenic centres of the amines had a strong influence on both the enantiomeric ratio and the reaction rate. Herein, molecular modelling approaches have been used to revise the lipase-catalysed aminolysis mechanism. Thus, complexes of CAL-B with the phosphonamidate analogues related to substrates in the kinetic resolution of several 2-phenylcycloalkanamines by this enzyme were built and minimised. This computational study suggests the formation of zwitterionic species (named TI-Z), resulting from the direct His-unassisted attack of the amine onto the carbonyl group of the acyl-enzyme, as the most plausible intermediate for the CAL-B-catalysed aminolysis. This proposal differs slightly from the commonly accepted serine-mediated mechanism, where removal of the proton from the amine occurs simultaneously to the nucleophile attack to the acyl-enzyme complex (TI-2). Subsequently, His-assisted deprotonation of the resulting ammonium group takes place, and a molecule of water could be necessary in some cases to facilitate the transfer of the proton to the catalytic histidine.  相似文献   

4.
Ultrasound was found to increase the rate of hydrolysis of a series of esters by up to 15%. No effect of molecular structure upon this enhancement was observed.  相似文献   

5.
6.
7.
8.
The mechanism of hydrolysis of n-nitrophenyl acetate (NPA), butyrate (NPB), caprylate (NPC), and o-methoxycinnamate (NPOMC) catalysed by benzyl-containing polyethyleneimines of linear and branched structures was investigated in aqueous media. The reaction seems to proceed via a general basic mechanism of catalysis and does not involve acylation of the catalyst. Benzyldiethylamine is an analogue of the active centres in polymers with pKa = 8 · 35 ± 0 · 1, localized in the polymer globules at sites of higher hydrophobity.The reaction has a three-step mechanism involving binding of the substrate to an active centre (to give Michaelis sorption complex), substrate conversion and desorption of products. For each step, rate constants were determined. The effect of polymer (K2/Km)/KII increases from NPA to NPC; in the latter case, it is of order 105.  相似文献   

9.
The reactions of the title substrate (1) with a series of secondary alicyclic amines are subjected to a kinetic investigation in 44 wt% ethanol‐water, at 25.0°C, ionic strength 0.2 M (KCl). Under amine excess over the substrate, pseudo‐first‐order rate coefficients (kobs) are obtained. Plots of kobs against [NH], where NH is the free amine, are nonlinear upwards, except the reactions of piperidine, which show linear plots. According to the kinetic results and the analysis of products, a reaction scheme is proposed with two tetrahedral intermediates, one zwitterionic (T±) and another anionic (T), with a kinetically significant proton transfer from T± to an amine to yield T (k3 step). By nonlinear least‐squares fitting of an equation derived from the scheme to the experimental points, the rate microcoefficients involved in the reactions are determined. Comparison of the kinetics of the title reactions with the linear kobs vs. [NH] plots found in the same aminolysis of O‐ethyl 4‐nitrophenyl dithiocarbonate (2) in the same solvent shows that the rate coefficient for leaving group expulsion from T± (k2) is larger for 2 due to a stronger push by EtO than PhO. The k3 value is the same for both reactions since both proton transfers are diffusion controlled. Comparison of the title reactions with the same aminolysis of phenyl 4‐nitrophenyl thionocarbonate (3) in water indicates that (i) the k2 value is larger for the aminolysis of 1 due to the less basic nucleofuge involved and the small solvent effect on k2, (ii) the k3 value is smaller for the reactions of 1 due to the more viscous solvent, (iii) the rate coefficient for amine expulsion from T± (k−1) is larger for the aminolysis of 1 than that of 3 due to a solvent effect, and (iv) the value of the rate coefficient for amine attack (k1) is smaller for the aminolysis of 1 in aqueous ethanol, which can be explained by a predominant solvent effect relative to the electron‐withdrawing effect from the nucleofuge. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 839–845, 1999  相似文献   

10.
The aminolysis of esters is a basic organic reaction considered as a model for the interaction of carbonyl group with nucleophiles. In the present computational study the different possible mechanistic pathways of the reaction are reinvestigated by applying higher level electronic structure theory, examining the general base catalysis by the nucleophile, and a more comprehensive study the solvent effect. Both the ab initio QCISD/6-31(d,p) method and density functional theory at the B3LYP/6-31G(d) level were employed to calculate the reaction pathways for the simplest model aminolysis reaction between methylformate and ammonia. Solvent effects were assessed by the PCM method. The results show that in the case of noncatalyzed aminolysis the addition/elimination stepwise mechanism involving two transition states and the concerted mechanism have very similar activation energies. However, in the case of catalyzed aminolysis by a second ammonia molecule the stepwise mechanism has a distinctly lower activation energy. All transition states in the catalyzed aminolysis are 10-17 kcal/mol lower than those for the uncatalyzed process.  相似文献   

11.
Immobilized Candida antarctica (Novozyme 435) catalyzed synthesis of N-acylethanolamines is described. Treatment of methyl esters with lipase and amines yielded the desired amides within 2-24 h with yields ranging from 41% to 98%.  相似文献   

12.
Plots of log k(0) vs pH for the cyclization of trifluoroethyl and phenyl 2-aminomethylbenzoate to phthalimidine at 30 degrees C in H(2)O are linear with slopes of 1.0 at pH >3. The values of the second-order rate constants k(OH) for apparent OH(-) catalysis in the cyclization reactions are 1.7 x 10(5) and 5.7 x 10(7) M(-)(1) s(-)(1), respectively. These rate constants are 10(5)- and 10(7)-fold greater than for alkaline hydrolysis of trifluoroethyl and phenyl benzoate. The k(OH) for cyclization of the methyl ester is 7.2 x 10(3) M(-)(1) s(-)(1). Bimolecular general base catalysis occurs in the intramolecular nucleophilic reactions of the neutral species. The value of the Bronsted coefficient beta for the trifluoroethyl ester is 0.7. The rate-limiting step in the general base catalyzed reaction involves proton transfer in concert with leaving group departure. The mechanism involving rate-determining proton transfer exemplified by the methyl ester in this series (beta = 1.0) can then be considered a limiting case of the concerted mechanism. General acid catalysis of the neutral species reaction or a kinetic equivalent also occurs when the leaving group is good (pK(a) 相似文献   

13.
14.
15.
The kinetics of aminolysis of 1-thio-β-D-glucopyranosyl esters of N-protected alanines (1) in dichloromethane, at 26°, by ethyl glycinate, under pseudo-first-order conditions follows the relationship k-obsd = k2[H-Gly-OEt]. Significant differences were observed in the rates of dipeptide forming aminolysis for 1-thiolesters 1a-1f, depending on N-acyl substituents and the configuration of alanine.  相似文献   

16.
17.
18.
Pyrrolidine was shown to deacetylate aromatic acetyl groups considerably faster than aliphatic groups.  相似文献   

19.
The reactions of S‐methyl O‐(4‐nitrophenyl) thiocarbonate ( 1 ) and S‐methyl O‐(2,4‐dinitrophenyl) thiocarbonate ( 2 ) with a series of secondary alicyclic (SA) amines and phenols are subjected to a kinetic investigation. Under nucleophile excess, pseudo‐first‐order rate coefficients (kobs) are obtained. Plots of kobs against the free nucleophile concentration at constant pH are linear with slopes kN. The Brønsted plots (log kN vs. nucleophile pKa) for the reactions are linear with slope (β) values in the 0.5–0.7 range, in accordance with concerted mechanisms. Comparison of the SA aminolysis of 1 with the same one carried out in water shows that the change of solvent from water to aqueous ethanol destabilizes the zwitterionic tetrahedral intermediate, changing the mechanism from stepwise to concerted. This destabilization is greater than that due to the change from SA amines to quinuclidines. For the phenolysis reactions, the kN values in aqueous ethanol are smaller than those for the same reactions in water. Considering that the nucleophile is an anion, this result is unexpected because the anion should be more stabilized in the more polar solvent. This result is explained by the facts that the phenoxide reactant has a negative charge that is delocalized in the aromatic ring and the transition state is highly polar. © 2011 Wiley Peiodicals, Inc. Int J Chem Kinet 43: 353–358, 2011  相似文献   

20.
《Tetrahedron: Asymmetry》1998,9(16):2777-2781
The bifunctional catalyst 2-hydroxypyridine (2-pyridone) is shown to promote the intramolecular aminolysis of the polyfunctionalized long-chain amino trichloroethyl ester 8 to afford cryptophycin-51 (4). This process for the construction of the macrocyclic core of cryptophycin derivatives is noteworthy for its convenience, avoidance of expensive coupling reagents, and use of mild reaction conditions.  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