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 171 毫秒
1.
Yiyu Ge 《Tetrahedron letters》2007,48(26):4585-4588
The formyl group was successfully removed from N-aryl formamide by KF on a solid support of basic Al2O3 in 4-20 min with microwave irradiation. The conditions mimic base-catalyzed hydrolysis of formamide and are compatible with carbamates and t-butyl esters, but not methyl, ethyl, and benzyl esters.  相似文献   

2.
In order to evaluate more precise kinetics parameters: rate constant k and Ea values for poly(l-lactic acid) hydrolysis, the reaction was carried out under high-pressure steam in a temperature range of 100-130 °C. Molecular weights of hydrolyzates were calculated by the universal calibration method without being influenced by any weight loss. The changes in molecular weight could be successfully explained according to the auto-catalytic hydrolysis mechanism, clearly indicating the critical point. Resulting k and Ea values were estimated as 8.4 × 10−5-7.2 × 10−4 s−1 and 87.2 kJ mol−1 with high R2 values, respectively. Moreover, to determine the deviation of the parameter values, influences of four factors on the measurements and calculation: (1) use of number-average molecular weight value alone, (2) use of relative molecular weight based on polystyrene standards, (3) weight loss during the hydrolysis, and (4) selection of reaction mechanism were evaluated quantitatively.  相似文献   

3.
The first representatives of chiral boron atom-containing amine-cyanomethoxycarbonyl boranes (A · BH(CN)COOMe) have been synthesized either from the corresponding amine-bromocyanomethoxycarbonylborane complexes with [Bu4N]CN or from Me3N · BH(CN)COOMe and an amine in a base-exchange reaction. Acid hydrolyses of methyl esters generated the free acids (A · BH(CN)COOH), which are isoelectronic to the α-cyano carboxylic acids. Their pKa values and hydrolysis half-lives in acidic medium (that is rate of proton reduction) have been determined. Similarly to the alpha cyano carboxylic acids, the cyano group attached to the boron (in alpha position to the COOH group) increased the acid strength of carboxy boranes with 2.0-2.5 orders of magnitude. Independently from the type of the amine, pKa values of the amine-cyanocarboxyboranes (6.34-5.82) decrease consistently with the increase of pKb values of the amines. Hydrolytic decomposition rate of the alkylamine complexes increases with increasing pKb values of the amines while the opposite was found for pyridine base complexes. When considering both types of the amines, hydrolysis half-lives of the complexes range over several orders of magnitude from 0.005 to 400 h. Based on these observations protonation of the amine nitrogen atom appears to be the rate determining step in the hydrolysis process. With loss of methanol, 2-NH2-py · BH(CN)COOMe transformed into a five membered lactam derivative. X-ray diffraction revealed that the pyridine ring is coplanar with the five membered lactam ring. In the crystal two molecules are connected in a head to tail arrangement by strong intermolecular H-bonds between N(2)-H and the carbonyl oxygen (O1) with a donor and acceptor distance of 2.867(3) Å. Three new cyanomethoxycarbonylborates having the composition of K[BHn(CN)3−nCOOMe] (n = 1, 2) and K[B(OH)(CN)2COOMe] have also been synthesized and their properties examined.  相似文献   

4.
Catalytic efficiencies of seven divalent metal acetylacetonate complexes [M(acac)2; M = Cd(II), Co(II), Cu(II), Fe(II), Ni(II), Pb(II), and Zn(II)] with respect to the water-crosslinking kinetics of vinyltrimethoxysilane-grafted ethylene-propylene copolymer (EPR-g-VTMS) were investigated to examine the effects of progressive changes in metal ion using ATR-FTIR spectroscopy. The hydrolysis activation energies of EPR-g-VTMS follows the order: No catalyst ≈ Ni(acac)2 > Co(acac)2 > Fe(acac)2 ≈ Zn(acac)2 > Cd(acac)2 ≈ Cu(acac)2 > Pb(acac)2. Interestingly, the kinetics results revealed that the plots of hydrolysis activation energies of EPR-g-VTMS containing M(acac)2 complexes and Eigen’s water exchange constants for corresponding metal ions showed a excellent linear relationship, suggesting that the reaction pathway for the silane water-crosslinking with hydrous M(acac)2 complex in EPR-g-VTMS system may be similar to that for water exchange of the metal ion in an aqueous system. Based on the knowledge of traditional kinetics studies by Eigen and Wilkins and hybrid sol-gel chemistry, the plausible catalytic mechanism for M(acac)2 complexes in EPR-g-VTMS system was proposed.  相似文献   

5.
In vitro degradation of poly(ethyl glyoxylate) (PEtG), a functionalised polyacetal, was investigated. First, the thermodynamic polymerization parameters and the ceiling temperature (Tc) were determined (ΔHp = 28 ± 3 kJ mol−1, ΔSp = 98 ± 7 J mol−1 K−1, Tc = 310 ± 4 K). Secondly, PEtG hydrolysis was investigated using potentiometry, weight loss measurements, SEC and 1H NMR. The results show that PEtG is stable for at least 7 days in aqueous media. Then degradation occurs and releases ethanol and glyoxylic acid hydrate as final products. A scheme for the degradation mechanism involving chain scission and ester hydrolysis is proposed.  相似文献   

6.
A mild protocol for transesterification of simple esters is described. The method is based on the use of t-BuNH2/ROH (R = Me, Et, i-Pr, t-Bu) with or without LiBr. The scope of the procedure was explored for aliphatic and aromatic esters. The protocol is particularly useful when going from higher to lower hindered esters and harsh reaction conditions are needed for the reversal process. A rationalization of the mechanism is presented. The scope and limitation of this transformation are also described.  相似文献   

7.
Isothermal titration calorimetry (ITC) has been used to observe the chitinase-catalyzed hydrolysis of tetra-N-acetylchitotetraose. Enzymatic hydrolysis of tetra-N-acetylchitotetraose by chitinase B from Serratia marcescens produces exclusively two molecules of di-N-acetylchitobiose allowing for the determination of a single glycosidic bond hydrolysis heat that was used to monitor the rate of the enzymatic reaction. The change in heat rate with respect to time (dQ/dt) was translated to the reaction rate, and the total heat produced was related to substrate concentration throughout the reaction. Reaction rates versus substrates concentration were fit to Michaelis-Menten plots, yielding a kcat of 40.9 ± 0.5 s−1 and a Km of 54 ± 2 μM.  相似文献   

8.
The activation of bovine liver arginase, which catalyzes the hydrolysis of l-arginine to l-ornithine and urea, by manganese ions was studied by thermokinetic methods at 37 °C in 40 mM sodium barbiturate-HCl buffer solution (pH 9.4). Full activation of arginase, by incubation with 0.1 mM Mn2+, resulted in increased of Vmax, and a higher sensitivity of the enzyme to product and l-lysine inhibition, with no change in the Km for arginine. Upon addition of 0.1 mM Mn2+ to the reaction, the inhibitory constants of product (KP) and l-lysine (KI) decreased from 1.18 to 0.70 mM and from 5.60 to 3.10 mM, respectively. We suggest that the exogenous manganese ions in reaction recovered the activity of arginase, which was lost in dissolving and dilution, without effecting on the mechanism of the reaction.  相似文献   

9.
A series of new asymmetrically N-substituted derivatives of the 1,4,7-triazacyclononane (tacn) macrocycle have been prepared from the common precursor 1,4,7-triazatricyclo[5.2.1.04,10]decane: 1-ethyl-4-isopropyl-1,4,7-triazacyclononane (L1), 1-isopropyl-4-propyl-1,4,7-triazacyclononane (L2), 1-(3-aminopropyl)-4-benzyl-7-isopropyl-1,4,7-triazacyclononane (L3), 1-benzyl-4-isopropyl-1,4,7-triazacyclononane (L4) and 1,4-bis(3-aminopropyl)-7-isopropyl-1,4,7-triazacyclononane (L5). The corresponding monomeric copper(II) complexes were synthesised and were found to be of composition: [Cu(L1)Cl2] · 1/2 H2O (C1), [Cu(L4)Cl2] · 4H2O (C2), [Cu(L3)(MeCN)](ClO4)2 (C3), [Cu(L5)](ClO4)2 · MeCN · NaClO4 (C4) and [Cu(L2)Cl2] · 1/2 H2O (C5). The X-ray crystal structures of each complex revealed a distorted square-pyramidal copper(II) geometry, with the nitrogen donors on the ligands occupying 3 (C1 and C2), 4 (C3) or 5 (C4) coordination sites on the Cu(II) centre. The metal complexes were tested for the ability to hydrolytically cleave phosphate esters at near physiological conditions, using the model phosphodiester, bis(p-nitrophenyl)phosphate (BNPP). The observed rate constants for BNPP cleavage followed the order kC1 ≈ kC2 > kC5 ? kC3 > kC4, confirming that tacn-type Cu(II) complexes efficiently accelerate phosphate ester hydrolysis by being able to bind phosphate esters and also form the nucleophile necessary to carry out intramolecular cleavage. Complexes C1 and C2, featuring asymmetrically disubstituted ligands, exhibited rate constants of the same order of magnitude as those reported for the Cu(II) complexes of symmetrically tri-N-alkylated tacn ligands (k ∼ 1.5 × 10−5 s−1).  相似文献   

10.
The mechanisms underlying the hydrolysis of methyl acetate and acetamide under acidic conditions were investigated using the MP2/6-311+G(d,p)//MP2/6-31+G(d,p) level of theory. It was necessary to include two water molecules as reactants to obtain a tetrahedral (TD) intermediate for the AAC2 mechanism that Ingold classified for the hydrolysis of methyl acetate. This mechanism includes two TS structures, one for the formation of the TD intermediate and the other for its decomposition. Since the activation energies were calculated to be 15.7 and 18.3 kcal mol−1, the second step determines the rate of hydrolysis. The calculated value was close to that observed at ∼16 kcal mol−1. It was confirmed that the AAC2 mechanism had a barrier lower by 9.9 kcal mol−1 than the AAL2 mechanism. The AAC2 mechanism is also applicable to the acid-catalyzed hydrolysis of acetamide. It is not the TD intermediate with which the NH3+ moiety forms, but one further step is required to produce the final products, acetic acid and ammonium ion.  相似文献   

11.
To mimic the phosphate ester hydrolysis behavior of purple acid phosphatases the heterobimetallic complex [(BNPP)FeIIIL(μ-BNPP)NiII(H2O)](ClO4) (1) has been synthesized from the precursor complexes [FeIII(LH2)(H2O)2](ClO4)3·3H2O and [FeIII(LH2)(H2O)Cl](ClO4)2·2H2O. In these compounds, L2− is the anion of the tetraiminodiphenol macrocyclic ligand (H2L), while LH2 is the zwitterionic form in which the phenolic protons are shifted to the two metal-uncoordinated imine nitrogens, and BNPP is bis(4-nitrophenyl)phosphate. The X-ray crystal structure of compound 1 has been determined. The structure of 1 comprises of two edge-shared distorted octahedrons whose metal centers are bridged by two equatorial phenolate oxygens and two axially disposed oxygens of a BNPP ligand. The internuclear Fe?Ni distance is 3.083 Å. The high-spin iron(III) and nickel(II) in 1 are antiferromagnetically coupled (J = −7.1 cm−1; H = −2JS1·S2) with S = 3/2 spin ground state. The phosphodiesterase activity of 1 has been studied in 70:30 H2O-(CH3)2SO medium with NaBNPP as the substrate. The reaction rates have been measured by varying pH (3-10), temperature (25-50 °C), and with different concentrations of the substrate and complex at a fixed pH and temperature. Treatment of the rate data, obtained at pH 6.0 and at 35 °C, by the Michaelis-Menten approach have provided the following parameters: KM = 3.6 × 10−4 M, Vmax = 1.83 × 10−7 M s−1, kcat = 9.15 × 10−3 s−1. As compared to the uncatalyzed hydrolysis rate of BNPP, the kcat value is 8.3 × 108 times higher, showing that 1 behaves as an excellent model for phosphate ester hydrolysis.  相似文献   

12.
This Letter describes an attractive and efficient method for Mg(OR)2-mediated lactide alcoholysis. The catalysts were generated in situ from di-n-butylmagnesium and ROH to prevent aggregation of Mg(OR)2. The reaction of ROH [R = Me, Et, RCO2(Me)CH] with lactide initially yielded the ring-opened product HO[CH(CH3)CO]nOR (n = 2 or 3). The complete consumption of lactide caused the reaction to proceed further, giving environmentally friendly lactic acid esters in excellent yields under ambient conditions.  相似文献   

13.
The cyclization of 3- or 4-pentyn-1-ol is catalysed by PdCl2 or trans-[PdCl2L2] (L = R-camphorimine; R = Ph; Pri; NMe2) complexes at room temperature affording heterocyclic compounds, respectively, 2-methyl-2-pent-3-ynyloxy-tetrahydrofuran or 2-methyl-2-pent-4-ynyloxy-tetrahydrofuran which subsequently add water to give selectively 5-(2-methyl-tetrahydrofuran-2-yloxy)-pentan-2-one from both starting materials. By hydrolysis 5-(2-methyl-tetrahydrofuran-2-yloxy)-pentan-2-one undergoes ring cleavage to form 5-hydroxy-2-pentanone. The catalytic activity and selectivity of complexes trans-[PdCl2L2] (L = R-camphorimine) depend on the characteristics of the R group (NMe2 > Pri > Ph). The catalytic activity of PdCl2 is comparable to that of trans-[PdCl2L2] (L = Ph-camphorimine) which is the less efficient catalyst.  相似文献   

14.
The reaction of (chloromethyl)cyclopropane 5 with lithium and a catalytic amount of DTBB (5 mol %) in the presence of different carbonyl compounds [Et2CO, n-Pr2CO, (c-C3H5)2CO, (CH2)5CO, PhCOMe, t-BuCHO, i-PrCHO, PhCHO] as electrophiles in THF at −78 °C leads, after hydrolysis with water, to the corresponding cyclopropyl alcohols 6. However, when the same starting material is lithiated using naphthalene as the arene catalyst in ether at 0 °C and then reacts with the same series of electrophiles, the final hydrolysis with water yields the corresponding unsaturated alcohols 7.  相似文献   

15.
Photoexcitation of a solution of anthracene-9-methanol derived esters at ∼386 nm in CH3CN/H2O (3:2 v/v) results in fluorescence emission in the 380-480 nm range, with quantum yields of fluorescence (Φf) in the 0.01-0.09 range and releases of the carboxylic acids in good chemical yields (43-100%), with quantum yields of photoreaction (ΦPR, i.e., the photodisappearance of the esters) in the 0.067-0.426 range.  相似文献   

16.
(Liquid + liquid) equilibrium (LLE) data for the {water + acetic acid + dibasic esters mixture (dimethyl adipate + dimethyl glutarate + dimethyl succinate)} system have been determined experimentally at T = (298.2, 308.2, and 318.2) K. Complete phase diagrams were obtained by determining solubility curve and tie-line data. The reliability of the experimental tie-line data was confirmed by using the Othmer-Tobias correlation. The UNIFAC model was used to predict the phase equilibrium in the system using the interaction parameters determined from experimental data between CH2, CH3COO, CH3, COOH, and H2O functional groups. Distribution coefficients and separation factors were compared with previous studies.  相似文献   

17.
18.
A variety of N-protected indolylmethylbromides are carbonylated using 5 mol % Pd(PPh3)2Cl2 under Stille conditions in the presence of an alcohol to afford the corresponding methyl/ethyl esters.  相似文献   

19.
N-(3-ferrocenyl-2-naphthoyl) dipeptide esters (5-7) and N-(6-ferrocenyl-2-naphthoyl) dipeptide esters (8-10) were prepared by coupling either 3-ferrocenylnaphthalene-2-carboxylic acid 2 or 6-ferrocenylnaphthalene-2-carboxylic acid 4 to the dipeptide ethyl esters GlyAla(OEt) (5, 8), AlaGly(OEt) (6, 9), and AlaAla(OEt) (7, 10) using the standard N-(3-dimethylaminopropyl)-N′-ethylcarbodiimide hydrochloride (EDC), 1-hydroxybenzotriazole (HOBt) protocol. All the compounds were fully characterized using a combination of 1H NMR, 13C NMR, DEPT-135 and 1H-13C COSY (HMQC) spectroscopy, electrospray ionization mass spectrometry (ESI-MS) and cyclic voltammetry (CV). In vitro, the cytotoxic effects of compounds 5-10 show improvements over the corresponding N-(ferrocenyl)benzoyl derivatives, with IC50 values against the H1299 lung cancer cells ranging from 1.2 μM to 8.0 μM. N-(6-ferrocenyl-2-naphthoyl)-glycine-l-alanine ethyl ester 8 was found to be the most active derivative of the naphthoyl series so far, displaying an IC50 value of 1.3 ± 0.1 μM. This value is slightly lower than that found for the clinically employed anti-cancer drug cisplatin (IC50 = 1.5 ± 0.1 μM against H1299).  相似文献   

20.
The ionic liquid, butylmethylimidazolium tetrafluoroborate ([Bmim][BF4]), was found to be superior to classical organic solvents for the metal catalyzed regio- and stereoselective aminohalogenation of cinnamic esters. The aminohalogenation reaction of cinnamic esters with p-TsNCl2 proceeded at a faster rate (within 12 h) in the presence of a reduced amount of catalyst (CuOTf, 6.0 mol %). Good yields (76-82%) and excellent regio- and stereoselectivity (one isomer) were achieved for eight examples.  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