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.
Combined on-line transmission FTIR spectroscopy and band-target entropy minimization (BTEM) analysis were employed in order to monitor and analyze the kinetics of the alkaline hydrolysis reaction of diethyl phthalate (DEP) in aqueous-ethanol solvent mixture. This reaction is irreversible and involves two consecutive steps with the formation of the observable mono-ion intermediate species. The pure component mid-FTIR spectra of the reactive species involved in this reaction, namely DEP, mono-ion intermediate and di-ion product were successfully reconstructed using BTEM. Their corresponding concentrations were also calculated and subsequently employed to derive the kinetic rate parameters. The effect of temperature and the solvent mixture compositions on these two consecutive reaction steps were also discussed. The temperature variation study showed that both reaction rate coefficients increased with temperature. Both rate coefficients were also affected by the solvent mixture compositions and reached minimum values at certain water-ethanol solvent composition (circa 60% (v/v)). This study shows the utility of combined on-line transmission FTIR spectroscopy and chemometric techniques for the present, rather complex, consecutive organic reaction. Moreover, the present type of approach could facilitate better understanding of a wide variety of organic reactions that are performed in aqueous and mixed aqueous-organic solvents.  相似文献   

2.
The rates and products of cleavage of methyl (2-chloro-4-nitrophenyl) phosphate (2) promoted by a dinuclear Zn(II) complex (3) of 1,3-bis-N,N'(1,5,9-triazacyclododecyl)propane along with 1 equiv of ethoxide were investigated in ethanol solution containing small amounts of water (8 mM or=1.6 x 10(17) times relative to the background hydroxide reaction, suggesting that complex 3 promotes the hydrolysis at least 1000 times more effectively than ethanolysis.  相似文献   

3.
A commercial lipase, “Lipolase T100”, was immobilised onto silica by means of physical adsorption. The silica-bound lipase was subsequently exposed to 1 vol. % glutaraldehyde (pentane-1,5-dial). The silica was loaded repeatedly with the Lipolase T100 in 0.05 M Tris buffer (pH 8.5) until saturation was achieved. During the 1st, 2nd, 3rd, 4th, and 5th cycles of loading of silica with the enzyme, the protein-binding on the silica achieved 51.73 %, 48.27 %, 26.92 %, 10.73 %, and 4.29 %, respectively. The synthesis of methyl salicylate (methyl 2-hydroxybenzoate) and linalyl ferulate (3,7-dimethylocta-1,6-dien-3-yl 4-hydroxy-3-methoxycinnamate) carried out at 45°C under shaking with mole ratios of 200 mM of acid and 500 mM alcohol in DMSO using 15 mg mL?1 of hyper-activated biocatalyst resulted in yield(s) of 77.2 % of methyl salicylate and 65.3 % of linalyl ferulate in the presence of molecular sieves. The hyper-activated biocatalyst was more efficient than the previously reported silica-bound lipase with minimum leaching of the enzyme from the reaction mixture. The K m and V max of the free (0.142 mM and 38.31 μmol min?1 mL?1, respectively) and silica-bound lipase (0.043 mM and 26.32 μmol min?1 mg?1, respectively) were determined for the hydrolysis of p-NPP. During repeated esterification studies using silica-bound lipase, yields of 50.1 % of methyl salicylate after the 5th cycle, and 53.9 % of linalyl ferulate after the 7th cycle of esterification were recorded. In the presence of molecular sieves (30 mg mL?1) in the reaction mixture, the maximum syntheses of methyl salicylate (77.2 %) and linalyl ferulate (65.3 %) were also observed. In a volumetric batch scale-up, when the reaction volume was increased to 50 mL, 44.9 % and 31.4 % yields of methyl salicylate and linalyl ferulate, respectively, were achieved.  相似文献   

4.
Kinetic study of hydroxide anion catalyzed hydrolysis of ethyl acrylate has been carried in ethanol-water (10–50% v/v) binary systems at the temperature range 30 ± 0.1, 35 ± 0.1, 40 ± 0.1, and 45 ± 0.1°C. Calculated specific rate constant values decreases with increasing proportion of ethanol at all temperatures. The observed retardation of a base catalyzed hydrolysis reaction is explained on the basis of fact that the formation of polarized transition state is disfavored with increase in % of ethanol. The relation between the change in dielectric constant due to variation in binary mixtures and change in specific rate constant are explained on the basis of electrostatic and non electrostatic contributions of solvent mixtures. The variation of ΔG*, ΔH*, ΔS* with solvent composition and the specific effect of water on the reaction rate kinetics are also discussed.  相似文献   

5.
Biphen(OPi-Pr) and (COD)PtCl2 give Biphen(OPi-Pr)PtCl2 which upon treating with ethyl Grignard forms Biphen(OPi-Pr)PtEt2. The thermal decomposition of Biphen(OPi-Pr)PtEt2 was investigated in the temperature range of 353-383 K. The clean and quantitative formation of the Pt(Ethene) adduct was observed. X-ray structures of a molecule in the solid state of all three reaction products and two further related complexes with phenyl fingers instead of i-Pr have been determined. For the complexes with i-Pr fingers a decisive deviation from a square plane is observed in contrast to the complexes with phenyl fingers. The P-Pt-P angle increases from about 95° in Biphen(OPi-Pr)PtCl2 to about 120° in Biphen(OPi-Pr)Pt(Ethene), forcing the bridging C-C single bond of the biphenyl fragment as near as 4.17 Å to the Pt center. No through-space coupling between the bridging C atoms and the Pt center could be observed in 13C NMR spectroscopy. No bond lengthening of the bridging C-C single bond in the biphenyl fragment was observed in Biphen(OPi-Pr)Pt(Ethene) in comparison to the precursor complexes. The thermal decomposition of Biphen(OPi-Pr)PtEt2 can be described by a first-order kinetic and the activation parameters were determined (temperature range: 353-383 K; ΔH = 173.8 ± 16.2 kJ/mol and ΔS = 104.7 ± 44.1 J/(mol K)). The reaction kinetics were also measured for perdeuterated ethyl groups yielding in a kinetic isotopic effect of 1.56 ± 0.14 which was almost temperature-independent. Selective deuteration at α and β position of the ethyl group, respectively, showed that β-H elimination takes place fast in comparison to the complete thermolysis. In the temperature range of 333-353 K only a scrambling of the deuterium atoms was found without further decomposition (temperature range: 333-353 K; ΔscramH = 76.1 ± 15.2 kJ/mol, ΔscramS = −80.7 ± 45.5 J/(mol K) for Biphen(OPi-Pr)PtEt2-d6). The ethene is not lost during the scrambling process. The scrambling process is connected with a primary KIE decisively larger than 1.56. Biphen(OPi-Pr)Pt(Ethene) exchanges the coordinated ethene with ethene in solution as proven by labeling experiments. Both a dissociative and an associative mechanism could be shown to take place as ethene exchange reaction by means of VT1H NMR spectroscopy via line shape analysis (temperature range: 333-373 K; ΔassH = 26.9 ± 29.6 kJ/mol, ΔassS = −148.0 ± 87.5 J/(mol K), ΔdissH = 86.0 ± 6.5 kJ/mol, ΔdissS = 5.4 ± 17.8 J/(mol K)). The Pt(0) complex formed during the dissociative loss of ethene activates several substrates among them: O2, H2, H2SiPh2 via Si-H activation, MeI presumably via forming a cationic methyl adduct and ethane via C-H activation but it was proven that the bridging C-C single bond of the biphenyl fragment is not even temporarily broken. The materials were characterized by means of 1H NMR, 13C NMR, 31P NMR, 195Pt NMR, EA, MS, IR, X-ray analysis and polarimetric measurement where necessary.  相似文献   

6.
The base hydrolysis of (αβS) (salicylato) (tetraethylenepentamine)cobalt(III) has been investigated in MeOH + water and DMSO + water media (0–70% (v/v) cosolvents) at 20.0 ? t°C ? 35.0 and I = 0.10 mol dm?3 (ClO4?). The phenoxide species [(tetren)CoO2CC6H4O]+ undergoes both OH?-independent and OH?-catalyzed hydrolysis via SN1ICB and SN1CB mechanism, respectively. The OH?-independent hydrolysis of the phenoxide species is catalyzed by both DMSO + water and MeOH + water media, the former exerting a much stronger rate accelerating effect than the latter. The OH?-catalyzed reaction is strongly accelerated by DMSO + water medium but insensitive to the composition of MeOH + water medium up to 40% (v/v) MeOH beyond which it was not detectable under the experimental conditions. Data analysis has been attempted on the basis of the solvent stabilizing and destabilizing effects on the initial state and transition state of the concerned reactions. The nonlinear variation of the activation parameters, ΔH and ΔS, with solvent compositions presumably indicates that the solvent structural effects mediate the energetics of solvation of the initial state and transition state of the concerned reactions. The linearity in ΔH vs. ΔS plot accomodating all data for k1 and k2 paths in DMSO + water and MeOH + water further suggests that the solvent effects on these parameters are mutually compensatory.  相似文献   

7.
Using density functional theory (DFT) at the B3LYP level with the 6-31G(d,p) basis set, a general equation is derived relating activation energy with the distance between the two reactive centers (rGM), and the hydrogen-bonding angle (αGM) in an intramolecular proton transfer process. The strong correlation between the values of rGM and αGM with the activation energy, ΔG, which reflects the experimental reaction rate, provides an excellent tool to predict reaction rate based on calculated geometrical parameters for a certain system (ΔH, ΔG vs rGM and αGM). The slope of the equation can be used as an indicator of the mode by which the two reacting centers orchestrate in an intramolecular process.  相似文献   

8.
In this work, an analytical procedure was developed to monitor the ethanolysis of degummed soybean oil (DSO) using Fourier-transformed mid-infrared spectroscopy (FTIR) and methods of multivariate analysis such as principal component analysis (PCA) and partial least squares regression (PLS). The triglycerides (reagents) and ethyl esters (products) involved in ethanolysis were shown to have similar FTIR spectra. However, when the FTIR spectra derived from seven standard mixtures of triolein and ethyl oleate were treated by PCA at the region that represents the CO stretching vibration of ester groups (1700-1800 cm−1), only two principal components (PC) were shown to capture 99.95% of the total spectral variance (92.37% for the former and 7.58% for the latter PC). This observation supported the development of a multivariate calibration model that was based on the PLS regression of the FTIR data. The prevision capability of this model was measured against 40 reaction aliquots whose ester content was previously determined by size exclusion chromatography. Only small discrepancies were observed when the two experimental data sets were treated by linear regression (R2=0.9837) and these deviations were attributed to the occurrence of non-modeled transient species in the reaction mixture (reaction intermediates), particularly at short reaction times. Therefore, the FTIR/PLS model was shown to be a fast and accurate method to predict reaction yields and to follow the in situ kinetics of soybean oil ethanolysis.  相似文献   

9.
The complex formation reactions of [Cu(NTP)(OH2)]4? (NTP?=?nitrilo-tris(methyl phosphonic acid)) with some selected bio-relevant ligands containing different functional groups, are investigated. Stoichiometry and stability constants for the complexes formed are reported. The results show that the ternary complexes are formed in a stepwise mechanism whereby NTP binds to copper(II), followed by coordination of amino acid, peptide or DNA. Copper(II) is found to form Cu(NTP)H n species with n?=?0, 1, 2 or 3. The concentration distribution of the various complex species has been evaluated. The kinetics of base hydrolysis of glycine methyl ester in the presence of copper(II)-NTP complex is studied in aqueous solution at different temperatures. It is proposed that the catalysis of GlyOMe ester occurs by attack of OH? ion on the uncoordinated carbonyl carbon atom of the ester group. Activation parameters for the base hydrolysis of the complex [Cu(NTP)NH2CH2CO2Me]4? are, ΔH±?=?9.5?±?0.3?kJ?mol?1 and ΔS±?=??179.3?±?0.9?J?K?1?mol?1. These show that catalysis is due to a substantial lowering of ΔH±.  相似文献   

10.
8-Quinolinol (HQ) reacts with [Pd(α-/β-NaiR)Cl2] [α-/β-NaiR = 1-alkyl-2-(naphthyl-α-/β-azo)imidazole] in acetonitrile (MeCN) solution to give [Pd(α-/β-NaiR)(Q)](ClO4). The products are characterized by spectroscopic techniques (FT-IR, UV–Vis, NMR). The reaction kinetics show a first order dependence of rate on each of the concentration of the metal complex and HQ. Addition of LiCl to the reaction retarded the rate of reaction and has proved the cleavage of the Pd–Cl bond as the rate-determining step. Thermodynamic parameters (ΔH° and ΔS°) are determined from variable temperature kinetic studies. The magnitude of the second order rate constant, k2, increases as in the order: Pd(NaiEt)Cl2 < Pd(NaiMe)Cl2 < Pd(NaiBz)Cl2 as well as Pd(β-NaiR)Cl2 < Pd(α-NaiR)Cl2.  相似文献   

11.
The kinetics of alkaline hydrolysis of 2‐chloro‐3,5‐dinitrobenzotrifluoride 1 and 1‐chloro‐2,4‐dinitrobenzene 2 were studied in various acetonitrile–water (AN–H2O) mixtures (10–90% w/w) at different temperatures. Thermodynamic parameters ΔH# and ΔS# show great variation, whereas ΔG# appears to vary little with the solvent composition presumably due to compensating variations. The results are discussed in terms of the solvent parameters such as preferential solvation, dielectric constant, polarity/polarizability, and hydrogen bond donor and acceptor parameters. It has been found that the factors controlling the reaction rates are the desolvation of OH?, the solvophobicity of the medium, and free water molecules in rich AN mixed solvent. The data showed that the solvatochromic parameters of (AN–H2O) mixed solvent are destroyed in the presence of excess OH?. © 2010 Wiley Periodicals, Inc. Int J Chem Kinet 42: 453–463, 2010  相似文献   

12.
Palladium(II)-catalyzed oxidative cyclization of methyl 2-allyl-4,5,6-tribromo-3-hydroxybenzoate 4 gave a mixture of methyl 5,6,7-tribromo-2-methyl-benzofuran-4-carboxylate 5 and methyl 6,7,8-tribromo-2H-1-benzopyran-5-carboxylate 6; the ratio of the two products varied between 71:29 and 24:76 depending on the reaction conditions. On the other hand, NBS or NIS mediated cyclization of 4 followed by treatment with NaOMe or DBU furnished benzofuran derivative 5 in good overall yield.  相似文献   

13.
Solution thermodynamic parameters of selected non-polar solutes have been determined in the nematic and isotropic fluid states of di(p-methoxyphenyl)-trans-cyclohexane-1,4-dicarboxylate. These states have been reported to exert no measurable differential in kinetic medium effects on the Claisen rearrangement1. Partial molar enthalpies ΔHsoln2 and entropies ΔSsoln2 of solution of a series of substituted benzenes in the nematogenic solvent, determined by the gas—liquid chromatographic method of Martire, 2–4, are reported. Changes in solute excess Gibbs free energy ΔGE2 over the nematic—isotropic transition of the solvent, corresponding to the changes in free energy of solution ΔΔGsoln2, have been calculated for the series. The results show the nematic and isotropic states of the medium to exhibit distinctly different solvent characteristics and suggest that Claisen reaction kinetics in the nematogenic solvent reflect compensating medium effects on the reactant and its activated state.  相似文献   

14.
Substitution reaction of the labile SMe2 ligand in the cyclometalated platinum(II) complexes of general formula [PtAr(ppy)(SMe2)], 1, in which ppy = deprotonated 2-phenylpyridyl and Ar = p-MeC6H4 or p-MeOC6H4, by several N or P donor reagents were studied; the N-donors, N, are pyridine (Py) and substituted pyridines, N = 4-MePy, Py, Py-d5, 2-MePy, 3-PhPy, 3,4-Me2Py, 4-tBuPy or 3-C(O)OMePy, and the P-donors, L, are phosphines or phosphites, L = P(OPh)3, P(O-iPr)3, PPh3, PPh2Me and L2 = Ph2PCH2PPh2, bis(diphenylphosphino)methane (dppm). The products were identified by multinuclear NMR studies as [PtAr(ppy)(N)], 2, or [PtAr(ppy)(L)], 3, respectively. Complexes 1 have a MLCT band in the visible region which was used to easily follow the kinetics of the ligand substitution reactions by UV-vis spectroscopy. Although the complexes 1 contain two cis Pt-C bonds, the substitution reactions followed a normal associative mechanism. The rates of reactions were depended on the concentration and the nature of the entering group. The ΔHS compensation plot gave a straight line suggesting the operation of the same mechanism for all entering nucleophiles.  相似文献   

15.
The acid-catalysed hydrolysis of sulphilimines of XC6H4(Me)SNTs and MePhSNSO2C6H4Y type has been studied by a kinetic method in moderately concentrated (1–6 M) aqueous H2SO4 and HClO4 solutions. The rate law: rate = kψ[sulphilimine] is valid for hydrolysis leading to sulphoxides and sulphonamides. The dependence of kψ on acidity, temperature and substituents X and Y has been measured and interpreted, ?X, ?Y and ΔS data (+ 1·19, + 1·00 and −18·7- - 22·6 e.u., resp) show that the nucleophilic attack of water on the positively polarized S(IV) atom of protonated sulphilimines can be regarded as the rate-determining step of the hydrolysis. From φ parameters (0·94−1·5) calculated for the hydrolysis of MePhSNTs it follows that water participates in the reaction as a nucleophile and proton-transfer agent.  相似文献   

16.
The dimer-monomer reactions were investigated for the system cis and transo,o'-azodioxytoluene-o-nitrosotoluene in acetonitrile solvent. For the reaction cis dimer-monomer the following thermodynamic and activation parameters have been derived: ΔH°=58.5±2.5 kJ mole?1, ΔS°=206.2±3.8 J mole?1 K?1, ΔH=63.6±3.3 kJ mole?1, ΔS=6.3±0.3 J mole?1 K?1. The corresponding values for the reaction trans dimer-monomer are: ΔH°=45.6±2.1 kJ mole?1, ΔS°=162.7±7.1 J mole?1 K?1, ΔH=80.8±2.9 kj mole?1, ΔS=-13.4±0.8 mole?1 K?1. There is no evidence of a direct cis-trans isomerization (i.e. a reaction not proceeding via the monomer). NMR and various perturbation techniques monitoring the visible absorption of the monomer were employed.  相似文献   

17.
Picolinic acid (picH) reacts with [Pd(α-/β-NaiR)Cl2] [α-/β-NaiR = 1-alkyl-2-(naphthyl-α-/β-azo)imidazole] in acetonitrile (MeCN) medium to give [Pd(α-/β-NaiR)(pic)](ClO4). The products are characterized by spectroscopic techniques (FT-IR, UV–Vis, NMR). The reaction kinetics show first order dependence of rate on each of the concentration of Pd(II) complex and picH. Addition of LiCl to the reaction decreases the rate of reaction and has proved the cleavage of Pd–Cl bond at the rate-determining step. Thermodynamic parameters (Δ and Δ) are determined from variable temperature kinetic studies. The magnitude of the second order rate constant, k2 increases as in the order: Pd(NaiEt)Cl2 < Pd(NaiMe)Cl2 <  Pd(NaiBz)Cl2 as well as Pd(β-NaiR)Cl2 <  Pd(α-NaiR)Cl2.  相似文献   

18.
Degradation via hydrolysis is among the main transformation pathways and particularly for N‐methylcarbamates. Carbamate pesticide hydrolysis is known to proceed through alkaline catalysis, with reaction of the hydroxide ion with the carbonyl function or with abstraction of hydrogen in the α position with respect to the carbonyl. This reaction leads to the formation of methylamine and corresponding phenol. In this respect, the reaction kinetics of 4‐bromo‐3,5‐dimethylphenyl N‐methylcarbamate (BDMC) hydrolysis have been investigated in alkaline solution using a spectrophotometric technique and reversed phase liquid chromatography. The kinetic constants were determined following a proposed pseudo–first‐order kinetic model. The positive activation entropy ΔS = +35.73 J mol−1 K−1 and the absence of general base catalysis indicated an unimolecular elimination conjugate base (E1cB) hydrolytic mechanism involving the formation of methyl isocyanate. This result was confirmed by the fact that BDMC fits well into brönsted and Hammett lines, obtained for a series of substituted N‐methylcarbamate whose decomposition in aqueous media was established to follow an E1cB mechanism.  相似文献   

19.
Bis-azo-cyanocyclo-pentane, -hexane and -heptane are prepared and the kinetics of their thermolysis in several solvents are studied by differential scanning microcalorimetry. Unlike activation enthalpy and entropy, activation free enthalpy does not depend on solvent. Moreover its variation with temperature is quite small in each kinetic study temperature range. Thus ΔG# is given as an intrinsic stability characteristic of azo-nitriles. “Kinetic compensation effect” between ΔS# and ΔH# is discussed.  相似文献   

20.
The X-ray structure of two cyclopalladated acetate bridged dimers has been solved, and they are shown to exhibit an open-book type core, analogous to those that have been previously reported. Variable temperature NMR studies on these molecules demonstrates that this core persists in chloroform solution, resulting in restrictions to the movement of ancillary parts of the molecule. The barrier to rotation of a pendant phenyl ring against this core has been measured: ΔH = 56 kJ mol−1, and ΔS = 0.  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