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 15 毫秒
1.
The O?H bond dissociation energy (D O?H) has been estimated for 20 substituted 3-pyridinols and a substituted 3-pyrimidinol from experimental kinetic data by the intersecting parabolas method using α-tocopherol and 4-methoxyphenol as reference compounds. The following D O?H values (kJ/mol) have been obtained: 363.7 for 3-pyridinol, 365.3 for 2-alkyl-3-pyridinols (five compounds), 358.8 for 2-alkyl-6-methyl-3-pyridinols (six compounds), 378.1 for 5-benzyl-3-pyridinol, 353.2 for 2,4,6-trimethyl-3-pyridinol, 340.9 for 2-benzyl-6-methoxy-3-pyridinol, 345.8 for 2,6-dimethoxy-5-benzyl-3-pyridinol, 381.7 for 2-ethyl-4-nitro-6-methyl-3-pyridinol, 376.8 for 2-isopropyl-4-nitro-6-methyl-3-pyridinol, 318.3 for 2,4-dimethyl-6-dimethylamino-3-pyridinol, 357.3 for mexidol, and 322.2 for 2,4-dimethyl-6-dimethylamino-3-pyrimidinol. The substituent effect on the O?H bond dissociation energy in 3-pyridinols is considered. The stabilization energies of pyridinoxyl and phenoxyl radicals are compared. The activation energies and rate constants have been calculated for a series of reactions of various radicals with 3-pyridinols.  相似文献   

2.
O? H bond dissociation enthalpies (BDE) for a variety of substituted phenols were calculated using a combined quantum chemical method. It is found that the calculated O? H BDE correlated well with the recommended values, except for ortho‐tert‐butyl substituted phenols. For the electron‐donating group substituted phenols the calculated O? H BDE are slightly higher than the recommended values, however, for the electron‐withdrawing group substituted phenols the calculated O? H BDE are slightly lower than the recommended values.  相似文献   

3.
The O?H bond dissociation energy (D O?H) has been determined for eight alkylseleno-substituted phenols, one alkyltelluro-substituted phenol, and one alkyltelluro-substituted pyridinol. D O?H has been estimated by the intersecting-parabolas method from kinetic data using five reference compounds: α-tocopherol (D O?H = 330.0 kJ/mol), 3,5-di-tert-butyl-4-methoxyphenol (D O?H = 347.6 kJ/mol), 4-methylphenol (D O?H = 361.6 kJ/mol), 2,6-di-tert-butyl-4-methylthiophenol (D O?H = 336.3 kJ/mol), and 2,6-di-ter-tbutyl-4-methylphenol (D O?H = 338.0 kJ/mol). The following D O?H values (kJ/mol) have been obtained: 335.9 for 2,5,7,8-tetramethyl-2-phytyl-6-hydroxy-3,4-dihydro-2H-1-benzoselenopyran, 342.6 for 2-methyl-5-hydroxy-2,3-dihydrobenzoselenophene, 333.5 for 2,4,6,7-tetramethyl-5-hydroxy-2,3-dihydrobenzoselenophene, 339.4 for 2-tert-butyl-4-methoxy-6-octylselenophenol, 357.9 for dodecyl 3-(4-hydroxyphenyl) propyl selenide, 348.5 for dodecyl 3-(3,5-dimethyl-4-hydroxyphenyl)propyl selenide, 350.9 for dodecyl 3-(3-tert-butyl-4-hydroxyphenyl)propyl selenide, 338.0 for dodecyl 3-(3,5-di-tert-butyl-4-hydroxyphenyl) propyl selenide, 343.0 for 2,6-di-tert-butyl-4-(tellurobutyl-4′-phenoxy)phenol, and 338.8 for 6-octyltelluro-3-pyridinol. The stabilization energies of phenoxyl radicals containing R substituents (X = O, S, Se, Te) have been compared.  相似文献   

4.
Using ion-selective electrode(s) (ISE) the activity coefficients of NaCl in the system NaCl–NH4Cl–H2O at 10, 25, and 40°C were measured by a computer-controlled automatic titration system. The ionic strength fractions of NH4Cl were 0.2, 0.4, 0.6, and 0.8, respectively. It was found that the influence of the NH4+ cation on the Na glass ISE was small enough to be neglected up to 3.0 mol-kg-1. The Pitzer equation was adopted to calculate the activity coefficients of NaCl in this system and compared with the experimental data. Comparison of results indicates that the Pitzer parameters correlated from solubility data are suitable for calculating the activity coefficients for this system within the saturated solutions.  相似文献   

5.
Quantum chemical calculations are used to estimate the equilibrium S–NO bond dissociation energies (BDEs) for seven S-nitro-N-acetyl-d,l-penicillamine dipeptides (SNAP-two peptides) in acetonitrile solution. These compounds were studied by employing the hybrid density functional theory (B3LYP, B3P86 and B3PW91) methods together with the 6-31G** basis set. The obtained results are compared with available experimental results. It is demonstrated that B3PW91 method is the best method to compute the bond dissociation energies of SNAP-two peptides. The substituent and solvent effects of the S–NO BDEs are further analyzed. The results show that S–NO BDE increases with the increment of isoelectric points of substituted groups. In addition, the S–NO BDE decreases due to the inclusion of solvent effects. Furthermore, SNAP-two peptides and the other NO-donors are compared.  相似文献   

6.
The experimental data for the liquid- and gas-phase reactions of atoms and radicals with organoelement compounds R n – 1E–H
where E = Ge, Sn, P, and Se, are analyzed within the framework of the parabolic model of radical abstraction reactions. The parameters characterizing the activation energies of such reactions involving H, O, and F atoms and , R , aryl (A ), R , and nitroxyl (Am ) radicals are determined. The activation energies for thermally neutral reactions E e , 0 are calculated. Reactions of a hydrogen atom with the H–element bond are characterized by the close E e , 0 (kJ/mol) values: 51.4 (GeH4), 52.8 (PH3), and 52.6 (SeH2). The E e , 0 values for the reactions of alkyl radicals with the Ge–H and Sn–H bonds are also close: E e , 0 (kJ/mol) = 62.7 (R"3GeH) and 63.2 (R"3SnH). Low E e , 0 values are typical of the reactions of alkoxy radicals (E e , 0 (kJ/mol) = 43.9 (GeH4), 46.2 (R"3GeH), 48.9 (R"3SnH), 43.8 (PH3) and oxygen atoms (E e , 0 (kJ/mol) = 41.0 (GeH4) and 47.3 (SeH2). Higher E e , 0 values are found for the reactions of peroxy radicals (E e , 0 (kJ/mol) = 62.8 (R"3GeH) and 60.6 (R"3SnH)) and nitroxyl radicals (E e , 0 (kJ/mol) = 81.3 (R"3GeH) and 77.4 (R"3SnH). The atomic radius of element E affects the activation energy of a thermally neutral reaction. The E–H bond dissociation energies for seven germanium and two tin compounds, as well as for five phosphites, are calculated from the kinetic data in terms of the parabolic model.  相似文献   

7.
The experiments were performed to investigate the degradation of microcystins in order to assess the effectiveness and feasibility of UV/H2O2 system for the disinfection of water polluted by microcystins. The influence factors such as H2O2, pH and UV light intensities were investigated respectively. Degradation of microcystin-RR (MC-RR) could be fitted by either the pseudo-first-order or second-order rate equations. This homogenous system could significantly enhance the degradation rate due to the synergetic effect between UV and H2O2. The degradation mainly followed the mechanism of direct photolysis and .OH oxidation reactions. Experimental results showed that 94.83% of MC-RR was removed under optimal experimental conditions and the UV/H2O2 system provided an alternative to promote the removal of microcystins in drinking water supplies.  相似文献   

8.
The dissociation constants of six common human and veterinary antibiotics, namely, erythromycin, roxithromycin, tilmicosin, oleandomycin, josamycin, and spiramycin in 15?%, 25?%, 40?% and 50?% (v/v) methanol?Cwater solvent mixtures were determined by UV/pH titration and correlated with the Kamlet and Taft solvatochromic parameters, ?? ?, ?? and ??. Kamlet and Taft??s general equation was reduced to two terms by combined factor analysis and target factor analysis in these mixtures: the independent term and polarity/polarizability ?? ?, which are solvatochromic parameters. The influence of methanol on the dissociation constants was investigated. Further, the quasi-lattice quasi-chemical (QLQC) model of preferential solvation has been applied to quantify the preferential solvation by water of electrolytes in methanol?Cwater mixtures.  相似文献   

9.
Isoparametricity was observed in reactions of 3,5-dinitrophenyloxiran with arenesulfonic acids YC6H4SO3H in mixtures of dioxane with the dimethyl ether of diethyleneglycol (1:1, v:v) at 265, 287, and 303 K: at the isoparametric point for temperature, the rate of the process of oxiran ring-opening did not depend on the structure of Y, whereas at the isoparametric point for the constant of substituent Y, the free energy of activation did not depend on the temperature.  相似文献   

10.
The effect of the functional group on the partial rate constants of the reactions of the tert-butylperoxy radical with CH bonds in the cyclohexyloxy moiety of cyclohexyl acetate at 333 K was studied by the Howard–Ingold method. The ester group activates the CH bond at the 1-position and deactivates the CH bonds at the 3-position and, especially, the 2-position. The reactivity of CH bonds at the 4-position is close to the reactivity of CH bonds in cyclohexane.  相似文献   

11.
An asymmetric unactivated alkene/C H bond difunctionalization reaction for the concomitant construction of C CF3 and C O bonds was realized by using a Cu/Brønsted acid cooperative catalytic system, thus providing facile access to valuable chiral CF3‐containing N,O‐aminals with excellent regio‐, chemo‐, and enantioselectivity. Mechanistic studies revealed that this reaction may proceed by an unprecedented 1,5‐hydride shift involving activation of unactivated alkenes and a radical trifluoromethylation to initiate subsequent enantioselective functionalization of C H bonds. Control experiments also suggested that chiral Brønsted acid plays multiple roles and not only controls the stereoselectivity but also increases the reaction rate through activation of Togni’s reagent.  相似文献   

12.
《Analytical letters》2012,45(9):1907-1915
Abstract

The electrochemical behavior of citalopram was studied by square‐wave and square‐wave adsorptive‐stripping voltammetry (SWAdSV). Citalopram can be reduced and accumulated at a mercury drop electrode, with a maximum peak current intensity being obtained at a potential of approximately ?1.25 V vs. AgCl/Ag, in an aqueous electrolyte solution of pH 12. A SWAdSV method has been developed for the determination of citalopram in pharmaceutical preparations. The method shows a linear range between 1.0×10?7 and 2.0×10?6 mol L?1 with a limit of detection of 5×10?8 mol L?1 for an accumulation time of 30 s. The precision of the method was evaluated by assessing the repeatability and intermediate precision, achieving good relative standard deviations in all cases (≤2.3%). The proposed method was applied to the determination of citalopram in five pharmaceutical products and the results obtained are in good agreement with the labeled values.  相似文献   

13.
14.
This article describes a gas chromatography–mass spectrometry (GC–MS) method for the determination of flurbiprofen in pharmaceutical preparations. The method is based on the derivatization of flurbiprofen with N-methyl-N-(trimethylsilyl)trifluoroacetamide (MSTFA). For GC–MS, electron ionization mode (EI = 70 eV) and selected ion monitoring (SIM) mode were used for quantitative analysis (m/z 180 for flurbiprofen). Calibration curve was linear between the concentration range of 0.25–5.0 μg/mL. Intra- and inter-day precision values for flurbiprofen were less than 3.64, and accuracy (relative error) was better than 2.67%. The mean recovery of flurbiprofen was 99.4% for pharmaceutical preparations. The limits of detection and quantification of flurbiprofen were 0.05 and 0.15 μg/mL, respectively. No interference was found from tablet excipients at the selected assay conditions. Also, the method was applied for the quality control of five commercial flurbiprofen dosage forms to quantify the drug and to check the formulation content uniformity.  相似文献   

15.
The energetics of the phenol O–H bond in methanol and the water O–H bond in liquid water were investigated by microsolvation modelling and statistical mechanics Monte Carlo simulations. The microsolvation approach was based on density functional theory calculations. Optimised structures for clusters of phenol and the phenoxy radical with one and two methanol molecules are reported. By analysing the differential solvation of phenol and the phenoxy radical in methanol, we predict that the phenol O–H homolytic bond dissociation enthalpy in solution is 24.3±11 kJ/mol above the gas-phase value. The analysis of the water O–H bond dissociation by microsolvation was based on optimised structures of OH–(H2O)1–6 and –(H2O)1–7 clusters. Microsolvation modelling and statistical mechanics simulations predict that the HO–H bond dissociation enthalpies in the gas phase and in liquid water are very similar. Our results stress the importance of estimating the differences between the solvation enthalpies of the radical species and the parent molecule and the limitations of local models based on microsolvation.Proceedings of the 11th International Congress of Quantum Chemistry satellite meeting in honor of Jean-Louis Rivail  相似文献   

16.
A bioanalytical method has been developed and validated for determination of drospirenone in human plasma. The analytical method consists in the extraction of plasma sample with dichloromethane, followed by determination of drospirenone by LC–MS–MS using levonorgestrel as internal standard. Separation was achieved on a Peerless cyano column with an isocratic mobile phase consisting of methanol and ammonium formate buffer. Protonated ions formed by a Turboionspray in the positive mode were used to detect analyte and IS. The MS–MS detection was by monitoring the fragmentation for drospirenone and for levonorgestrel on a triple quadrupole mass spectrometer. The assay was calibrated over the range 5–100 ng mL?1 with a correlation coefficient of 0.9998. Validation data showed intra-batch (n = 6) CV% ≤ 5.58 and RE (%) between ?3.34 and +6.27 and inter-batch (n = 18) CV% < 6.08 and RE (%) between ?1.84 and +6.73. Mean extraction recovery were 83.31–92.58% for three QC samples and 89.32% for IS. Plasma samples were stable for three freeze-thaw cycles, or 24 h ambient storage, or 1 and 3 months storage at ?20 °C. Processed samples (ready for injection) were stable up to 72 h at autosampler (4 °C). This method has been used for analyzing plasma samples from a bioequivalence study with 12 volunteers.  相似文献   

17.
18.
Russian Journal of Physical Chemistry A - Model molecular systems structurally similar to the transition state of the limiting step of the hydrolysis of cephalosporin antibiotics by the L1...  相似文献   

19.
A simple and rapid LC–MS–MS assay was developed and validated for the quantitative determination of pitavastatin in human plasma. Sample pretreatment involved simple protein precipitation by addition of acetonitrile. Separation was on an Agilent 1.8 μm Zorbax SB-C18 column (150 mm × 4.6 mm) at 25 °C using isocratic elution with methanol–0.1% formic acid in water (85:15, v/v) at a flow rate of 0.4 mL min?1. Detection was performed using electrospray ionization in positive ion multiple reaction monitoring mode by monitoring the ion transitions m/z 422.0 → 290.1 for pitavastatin, and m/z 330.1 → 192.1 for paroxetine (IS). LC–MS–MS was found to improve the quantitation of pitavastatin in plasma and was successfully applied in pharmacokinetic studies.  相似文献   

20.
A method suitable for the determination of 3-nitrotyrosine, a well known marker of nitro-oxidative stress, in plasma by GC–ECD has been investigated. After deproteinisation, SPE and LC separation, 3-nitrotyrosine was derivatised by a single-step procedure to the corresponding N-heptafluorobutyryl heptafluorobutyl ester and analysed. The precision and reliability of the procedure were assured by using 4-nitrophenylalanine as internal standard and a diverter valve to protect the detector. The method was applied to the analysis of volunteers’ plasma samples.  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