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 24 毫秒
1.
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   

2.
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   

3.
The kinetics of alkaline hydrolysis of 2-chloroquinoxaline (QCl) with hydroxide ion was investigated spectrophotometrically at different percentages of aqueous–organic solvent mixtures with acetonitrile (10–60% v/v) and with dimethylesulphoxide (10–80%) over the temperature range from 25 to 45 °C. The reaction was performed under pseudo first order conditions with respect to 2-chloroquinoxaline (QCl). An increase in the percentage of organic solvent (v/v) has different effects on the reaction rate constants, presumably due to hydrogen bond donor and acceptor differences of the media and other solvatochromic parameters. The data were discussed in terms of the Kamelt-Taft parameter and E T(30). A nonlinear relation between the logarithm of the rate constant and reciprocal of the dielectric constant suggests the presence of selective solvation by the polar water molecules. Activation parameters ΔH #, ΔS # and ΔG # were determined and discussed.  相似文献   

4.
The dissociation energy of the O–H bond has been calculated by the homodesmotic reaction method for phenolic compounds, which are well-known antioxidants, including for natural phenols. Use of moderately complex computational levels, such as B3LYP/6-31G(d), is sufficient for reliably estimating the D(O–H) value for phenols within the homodesmotic approach. The O–H bond dissociation energy for monosubstituted phenols has been calculated, and the additive character of the effect of methyl groups on D(O–H) in methylphenols has been demonstrated: the introduction of a CH3 group into the aromatic ring decreases the D value by 7.8 kJ/mol (ortho position), 1.8 kJ/mol (meta position), and 7.6 kJ/mol (para position). The O–H bond strength has been calculated for a number of ubiquinols, selenophens, flavonoids, and chromanols. The D(O–H) value recommended for α-tocopherol is 328.0 ± 1.3 kJ/mol.  相似文献   

5.
The work deals with the establishment of the dependence of the vibrational frequencies of strong O–H?O and N–H?O hydrogen bonds for the diagnosing the bonds themselves. To this end, the Raman spectra of a large number of different normal and deutero-substituted crystals characterized by the presence of strong O–H?O and N–H?O bonds are measured and the quantum chemical calculation is performed for one of these compounds. The dependence of the O–H stretching frequency on the O?O distance is constructed differing from that previously known for short O?O contacts. The mechanisms of significant broadening of the O–H vibration band in strong O–H?O hydrogen bonds are considered. Different dependences of the N–H vibrational frequencies in N–H?O bonds are reported and the causes of this diversity are discussed.  相似文献   

6.
The mixture {yNaCl + (1 – y)CaCl2}(aq) has been studied with the hygrometric method at the temperature 25°C. The water activities were measured at total molalities from 0.25 mol-kg–1 to near saturation for different ionic-strength fractions y of NaCl with y = 0.33, 0.50, and 0.67. The obtained data allow the calculation of osmotic coefficients. The experimental results are compared with the predictions of the Zdanovskii–Stokes–Robinson (ZSR), Kusik and Meissner (KM), Robinson and Stokes (RS), Lietzke and Stoughton (LS II), Reilly–Wood–Robinson (RWR), and Pitzer models. From the measured osmotic coefficients, the Pitzer ionic mixing parameters Na Ca and Na Ca Cl are determined and are used to predict the solute activity coefficients in the mixtures. The excess Gibbs energy is also determined. These results are compared with those given in the literature.  相似文献   

7.
The geometries of the two H–O–C rotamers of synand two of anti-7-norbornenol have been optimized at the ab initio HF/6-31G** and B3LYP hybrid HF-DFT levels of theory by using a 6-31G** basis set. Contrary to an earlier report, we find that the (nongeometry constrained) anti-trans isomer (1d) is predicted to be more stable than the corresponding syn-cis form (1a). The increased stability of 1d vis-à-vis 1a can be accounted for in terms of relative H(1)–C(2)–O(3)–H(4) torsion energy effects. The computational results indicate that the hydroxyl proton in 1a enters into intramolecular hydrogen bonding with the proximate C=C bond. Supporting evidence for this conclusion resides in the fact that the 1a is predicted to possess the lowest O–H stretching frequency, a result that can be attributed to -hydrogen bonding.  相似文献   

8.
9.
《Fluid Phase Equilibria》2004,224(2):251-256
In this work, experimental liquid–liquid equilibria (LLE) of the limonene + ethanol + water system are presented. The LLE of this system has been measured at 293.15, 303.15, 313.15 and 323.15 K. The equilibrium data presented are correlated using NRTL and UNIQUAC equations. Finally, the reliability of these models is tested by comparison with experimental results.  相似文献   

10.
《Fluid Phase Equilibria》2002,202(2):221-231
The mixed aqueous electrolyte system magnesium and manganese sulfate has been studied with the hygrometric method at the temperature 298.15 K. The relative humidity of this system is measured at total molalities from 0.2 mol kg−1 to about saturation of one of the solutes for different ionic-strength fractions y of MgSO4 with y=0.2, 0.5 and 0.8. The obtained data allow the deduction of new thermodynamic parameters. The experimental results are compared with the predictions of ZSR rule. From these measurements, the new Pitzer mixing ionic parameters are determined and used to predict the solute activity coefficients in the mixture. The obtained results are used to calculate the excess Gibbs energy at total molalities for different ionic-strength fractions y of MgSO4.  相似文献   

11.
Glycoproteins are an important class of proteins that play a significant role in many cellular events. In the present study, we analyze the influence of C–H…O interactions in relation to other environmental preferences in glycoproteins. CH…O interactions are now accepted as a genuine hydrogen bond. Main chain–main chain interactions are predominant. Proline residues stabilize strands by C–H…O interactions in glycoproteins. Majority of the C–H…O interacting residues were conserved and had one or more stabilization centers. CH…O interactions might be responsible for the global conformational stability, since long-range CH…O contacts were predominant. The results presented in this study might be useful for structural stability studies in glycoproteins.  相似文献   

12.
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   

13.
Suqin Han  Erbao Liu  Hua Li 《Mikrochimica acta》2005,149(3-4):281-286
A flow injection chemiluminescence method has been developed and applied to the determination of hemin in tablets and animal blood. The proposed method is based on the luminescent properties of the Rhodamine B–H2O2–NaOH system and the addition of sodium dodecyl sulfonate (SDS) as emission-sensitizer. Hemin was determined over the concentration range of 8.6×10–10–8.6×10–7M with a detection limit of 8.6×10–11M (3). The relative standard deviation (RSD) for seven independent detections of 1.72×10–8M hemin was 3.0%. The proposed method was successfully applied to the analysis of hemin in pharmaceutical preparations and animal blood with a recovery of 96–108%. A possible CL mechanism of the present system was discussed, and free radicals were suggested to be involved in this reaction.  相似文献   

14.
ESR andIR spectroscopy and quantum chemical calculations are used to obtain the mechanistic data on the reaction . The IR bands are assigned by simulating the vibrational spectra of model low-molecular compounds. Quantum chemical calculations provided the data on the shapes of potential energy surface for the systems under study and transition states. These data are used to interpret the experimental data. The title reaction occurs much more readily in the case of organosilicon peroxy radicals than in the case of their hydrocarbon analogs. Surface silanone groups of silica react with ethylene molecules to form the siloxacyclobutane group.  相似文献   

15.
The O–H bond dissociation enthalpy (BDE) of 3,5-dimethylphenol has been determined, by using the EPR radical equilibration technique, as 362.5 ± 1.5 kJ/mol. This value is 7 kJ/mol smaller than that of phenol, this indicating that alkyl substitution of the meta hydrogens of phenol induces a weakening of the O–H bond, in contrast with what was reported in a recent calorimetric study.  相似文献   

16.
Development of rhodium catalysed O–H insertion reactions employing α-diazophosphonates with appropriately protected adenosine, uridine and thymidine derivatives is described. This synthetic methodology leads, following deprotection, to novel phosphononucleoside derivatives bearing a carboxylic acid moiety adjacent to the phosphonate. Protection strategies are critical to the success of the key O–H insertion. There are two important aspects: avoiding competing insertion pathways or catalyst poisoning, and being able to achieve deprotection without degradation of the phosphononucleosides.  相似文献   

17.
Russian Journal of Physical Chemistry A - Isothermal phase diagrams of ternary systems fullerenol-d–LaCl3–H2O and fullerenol-d–GdCl3–H2O at 25°C are studied via...  相似文献   

18.
The kinetics of decomposition of hydrogen peroxide in the liquid phase of the ternary system LiOH-H2O2-H2O was studied in the presence the solid phase of Li2O2·H2O and without it. The main kinetic parameters of the processes studied were determined.  相似文献   

19.
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   

20.
《Fluid Phase Equilibria》2004,216(2):229-233
The water activity and osmotic coefficients of the system {y NH4NO3+(1-y) KNO3}(aq) has been measured at total molalities from 0.2 mol kg−1 to about saturation of one of the solutes for different ionic-strength fractions y of NH4NO3 with y=0.2, 0.5 and 0.8 at the temperature 298.15 K using the hygrometric method. The obtained data allow the deduction of the thermodynamic parameters. From these measurements, new Pitzer ionic mixing parameters are determined and used to predict the solute activity coefficients in the mixture. The results obtained are used to calculate the excess Gibbs energy at total molalities for different ionic-strength fractions of NH4NO3.  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