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 359 毫秒
1.
Six voltaic cells have been evaluated for their suitability in the determination of thermodynamic parameters. The cells were prepared with all species present at conditions that approximate standard-state conditions and cell potentials were measured as a function of temperature. From these measurements graphs of voltage versus temperature were prepared. From these graphs it was possible to determine the standard heats of reaction (IIrxn), standard change in entropy or disorder (S°rxn), and the Gibbs free energy (G°rxn) for the spontaneous oxidation-reduction reactions. The standard cell potential (E°cell values were also calculated. Two cells with opposite temperature dependence of the cell potential were found to produce good agreement with the literature values of G°rxn, H°rxn, S°rxn and E°cell. Criteria for identifying additional cells that may be suitable for potentiometric studies of thermodynamic parameters are also included.  相似文献   

2.
Dilution enthalpies, measured using isothermal flow calorimetry, are reported for aqueous solutions of BaCl2 at 300°C and 11.0 MPa, MgCl2, CaCl2, and BaCl2 at 325°C and 14.8 MPa, and at 350°C and 17.6 MPa. Previously collected dilution enthalpies for aqueous solutions of MgCl2 and CaCl2 at 300°C and 10.3 MPa and for aqueous solutions of HCl at 250, 275, and 300°C at 10.3 MPa and 320°C at 12.8 MPa were included with the new data at 300°C and 11.0 MPa and at 350°C and 17.6 MPa when fitting the Pitzer parameters. The concentration range of the chloride solutions was 0.5 to 0.02 molal. Parameters for the Pitzer excess Gibbs ion–interaction equation were determined from the fits of the experimental heat data. Equilibrium constants, enthalpy changes, entropy changes, and heat-capacity changes for the association of alkaline earth metal ions and H+ with chloride ion were estimated from the heat data. For all systems, the enthalpy and entropy changes are positive and show accelerating increases with temperature. The resulting equilibrium constants show significant, but smaller, increases with temperature.  相似文献   

3.
The interaction of adenosine 5-monophosphate (AMP), adenosine 5-diphosphate (ADP), and adenosine 5-triphosphate (ATP) ions with protons in aqueous solution has been studied calorimetrically from 50 to 125°C and 1.52 MPa. At each temperature, the reaction of acidic AMP with tetramethylammonium hydroxide was combined with the heat of ionization for water to obtain the enthalpy of protonation of AMP, while the reactions of HCl with deprotonated tetramethylammonium salts of ADP and ATP were used to obtain the enthalpies of protonation of ADP and ATP. Equilibrium constant K, enthalpy change Ho, entropy change So, and heat capacity change C p o values were calculated for the stepwise protonation reactions as a function of temperature. The reactions involving the first protonation of AMP, ADP, and ATP and the third protonation of ADP and ATP were endothermic over the temperature range studied, while that involving the second protonation is exothermic for AMP and ADP, but is exothermic below 100°C and endothermic at 125°C in the case of ATP. Consequently, log K values for the first and third protonation reactions (phosphate groups) increase while those for the second protonation reaction (N1-adenine) decrease in the cases of AMP and ADP and go through a minimum in the case of ATP as temperature increases. The Ho values for all protonation reactions increase with temperature. The magnitude and the trend for the Ho, So, and C p o values with temperature are discussed in terms of solvent-solute interactions. The magnitude of the C p o values for the second protonation is consistent with little interaction between the phosphate ion and the protonated N1 site of the adenine moiety in AMP, but indicates moderate interaction between these groups in ADP, and strong interaction in ATP.  相似文献   

4.
5.
Conductance data for dilute solutions of S-n-butylisothiouronium bromide and iodide salts as well as S-alkylisothiouronium picrates (alkyl=methyl to octyl) are reported in pure nitrobenzene and methanol at 25, 35, and 45°C. The data were analyzed using Fuoss' equation (1980) (F-80) to obtain the three adjustable parameters; molar conductance at infinite dilution o, association constant KA and the association distance R that minimize the standard deviation . The variation of the parameters for each solvent are discussed with the variation of both cationic and anionic size as well as with temperature. It was found that the solvent-separated ion pair model (SSIP) can be applied in methanol for the salts under investigation. The thermodynamic parameters related to ion-pair formation were calculated and interpreted. The plots of #x039B;o vs. the reciprocal of the molecular weight M were used to separate the molar conductances into values for the individual ions.  相似文献   

6.
Experimental viscosities provide information on the structure of liquids and are required in the design of processes, which involve fluid flow, mass transfer, or heat transfer calculations. This work reports experimental viscosity data of the binary mixtures: 1-decanol + hexane, 1-decanol + diethylamine, and hexane + diethylamine at 10, 25, and 40°C and atmospheric pressure for the whole range of compositions. The viscosities of the pure liquids and their mixtures were determined using Cannon Fenske viscometers thermostated at ±0.01°C. The estimated error in the measured viscosities was less than ±0.005 mPa-s. The dynamic viscosity and the excess energy of activation for viscous flow were also calculated. The equation of Redlich–Kister was used for fitting the excess properties of the binary mixtures. The excess viscosity shows positive deviations from ideal behavior for the mixtures 1- decanol + hexane and 1-decanol + diethylamine and a small negative deviation for the binary system hexane + diethylamine. The experimental results have been also used to test some empirical and semiempirical equations adopted previously to correlate viscosity composition data.  相似文献   

7.
Viscosity and density data are reported for n-decane + propyl ethanoate, propyl propanoate, propyl butyrate, and n-tetradecane + propyl ethanoate, propyl propanoate, and propyl butyrate at 25°C and atmospheric pressure. Kinematic viscosities were determined using a capillary viscosimeter and densities were measured using vibrating-tube densimetry. The equations of Grunberg–Nissan, McAllister, Auslander, and Teja were fitted to the viscosity data. Excess molar Gibbs free energies of activation for flow were also evaluated. The experimental values obtained for excess volumes were compared with the Nitta et al. group contribution model.  相似文献   

8.
《Fluid Phase Equilibria》1999,157(2):317-342
Kinematic viscosities (ν) of the ternary ethane-1,2-diol (1)+2-methoxyethanol (2)+water (3) solvent system have been measured for 36 ternary mixtures covering the whole miscibility range expressed by the condition 0<X1,X2,X3<1, at 19 temperatures in the range −10≤t (°C)≤80. The measured values have been used to test some empirical equations of the type ν=ν(t) and ν=ν(Xi), in order to provide for useful interpolation procedures to obtain calculated values in correspondence to the experimental data gaps. From the experimental data, the excess kinematic viscosities (νE) have been calculated. Sign and magnitude of these quantities have been discussed in terms of type and nature of specific intermolecular interactions. Furthermore, derived quantities such as thermodynamic parameters of the viscous flow (ΔG*, ΔH* and ΔS*), have been analysed on the basis of the Eyring's model. All the investigated excess mixing properties indicate the probable absence of stable three-component adducts in this ternary solvent system.  相似文献   

9.
The viscosities of solutions of tetrapropylammonium bromide (Pr4NBr), tetrabutylammonium bromide (Bu4NBr), tetrapentylammonium bromide (Pen4NBr), tetrahexylammonium bromide (Hex4NBr), tetraheptylammonium bromide (Hep4NBr), tetraoctylammonium bromide (Oct4NBr), tetrabutylammonium tetraphenylborate (Bu4NBPh4), sodium tetraphenylborate (NaBPh4), and potassium tetraphenylborate (KBPh4) in N,N-dimethylacetamide are reported at 25°C. The viscosity data havebeen analyzed by the Jones-Dole equation for associated electrolytes to evaluate the viscosity B coefficients of the electrolytes. These data have also been analyzed by the transition-state theory to obtain the contribution of the solutes to the free energy of activation for viscous flow of the solution. The ionic contribution to the viscosity B coefficient and the free energy of activation for viscous flow have been estimated using of the reference electrolyte Bu4NBPh4. The bromide, tetraphenylborate, and tetraalkylammonium ions are found to be weakly solvated in N,N-dimethylacetamide, whereas significant solvation has been detected for sodium and potassium ions. The viscosity of the solvent is greatly modified by the presence of all the ions investigated here with the exception of the bromide ion.  相似文献   

10.
The oxidation wave of iodide in 0.075 mol-L–1 H2SO4 was analyzed at 25, 40, 55, 70, and 85°C. The reversibility of the I2/I system was checked using logarithmic transforms, half-wave potentials, and by studying I –1 = f(–1/2). The limiting currents obtained enabled us to determine the diffusion coefficient of I using Newman's equation. These experimental results were compared with Nernst's limiting values. The Stokes–Einstein equation is not verified. Hydration numbers for I at different temperatures were established. An empirical equation is proposed to predict the evolution of diffusion coefficients in a sulfuric acid medium with temperature.  相似文献   

11.
Conductivity measurements of oxalic acid and neutral oxalates (disodium oxalate, dipotassium oxalate, dicesium, and diammonium oxalate) were performed on dilute aqueous solutions, c < 3 × 10–3 mol-dm–3, from 5 to 35°C. These data and those available from the literature were analyzed in terms of dissociation steps of oxalic acid, the Onsager conductivity equation for neutral oxalates, the Quint–Viallard conductivity equation for the acid, and the Debye–Hückel equation for activity coefficients, to give the limiting equivalent conductances of bioxalate anion ;(HC2O4 ) and oxalate anion (1/2C2O4 2–) and the corresponding dissociation constants K 1 and K 2.  相似文献   

12.
Excess molar volumes for binary mixtures of acetonitrile + dichloromethane, acetonitrile + trichloromethane, and acetonitrile + tetracloromethane at 25°C have been used to calculate partial molar volumes , excess partial molar volumes , and apparent molar volumes of each component as a function of composition. The V m Evalues are negative over the entire composition range for the systems studied. The applicability of the Prigogine–Flory–Patterson theory was explored. The agreement between theoretical and experimental results is satisfactory for the systems with dichloromethane and tetrachloromethane. For the unsymmetrical behavior of the system with trichloromethane, however, the agreement is poor.  相似文献   

13.
ΔE T N values for 2-pyrrolidinone and N-methylbenzenesulfonamide solvent systems, in which the solvents were benzyl alcohol, 1,4-dioxane and hexamethylphosphoric triamide, were determined over the whole mole fraction range. The study was carried out at 30 and 50°C. The ΔE T N values were positive for all of these systems, with the exception of the 2-pyrrolidinone-hexamethylphosphoric triamide system, which was slightly negative. The results are discussed in terms of intermolecular interactions and preferential solvation.  相似文献   

14.
Experimental data of densities and viscosities are presented for the system 4-methyl-N-butylpyridinium tetrafluoroborate + methanol at 25, 40, and 50, 323.15 K and ambient pressure using a vibratage-tube densimeter and an Ubbelohde viscometer. Excess molar volumes V E and excess logarithm viscosities (ln )E have been determined. V E is negative and (ln )E positive over the entire mixture composition.  相似文献   

15.
Solubility of light fullerenes (C60, C70, and the standard fullerene mixture containing (wt %): C60 65, C70 34, C n>70 1) in the oleic, linoleic and linolenic acids, respectively, at 20–80°C was studied and the corresponding solubility polytherms were reported.  相似文献   

16.
The osmotic coefficients of lithium chloride, lithium bromide, and lithium nitrate in 2-propanol have been measured by the isopiestic method at 25°C. Sodium iodide was used as the isopiestic standard. The molality ranges covered were from 0.2 to 1.5 for LiCl and LiBr, and to 1.9 mol-kg-1 for LiNO3. The system of equations developed by Clegg–Pitzer and Pitzer were used to fit each set of osmotic coefficients. The experimental osmotic coefficient data are successfully correlated with these models. The parameters from the fit were used to calculate the mean molal activity coefficients.  相似文献   

17.
Present paper reports the measured densities (ρ) and refractive indices (n D) of aqueous solutions of ceric ammonium nitrate (CAN) at 20, 25, 30, and 35°C in different concentrations of solution. Apparent molar volumes (φv) have been calculated from the density data at different temperatures and fitted to Massons relation to get limiting partial molar volumes (? v 0 ) of CAN. Refractive index data were fitted to linear dependence over concentration of solutions and values of constant K and n D 0 for different temperatures were evaluated. Specific refractions (R D) of solutions were calculated from the refractive index and density data. Concentration and temperature effects on experimental and derived properties have been discussed in terms of structural interactions.  相似文献   

18.
In order to study the influence of aluminium on the oxidation resistance of titanium at high temperature, a range of binary alloys containing 1.65, 3, 5 and 10 wt.% of aluminium was prepared. Their oxidation kinetics were studied at temperatures between 500 and 750 °C using either continuous thermogravimetry or daily weighing for periods of up to several thousand hours. The results obtained for oxidation in air confirm the beneficial role of aluminium which has been observed previously for oxidation in oxygen. With regard to morphology and structure, aluminium modifies the internal structure of the oxide layers and their growth laws. A general dispersion of aluminium in rutile is observed, although a concentration of this phase is noted near the external interface; there is also a reduction in the amount of oxygen dissolved in the metal substrate which is related to the aluminium content in the alloy. Moreover, the presence of aluminium also modifies the adhesion of the oxide layers to the substrate.  相似文献   

19.
Excess molar volumes VmE at 25°C and atmospheric pressure over the entire composition range for binary mixtures of 1-hexanol with n-polyethers: 2,5-dioxahexane, 3,6-dioxaoctane, 2,5,8-trioxanonane, 3,6,9-trioxaundecane, 5,8,11-trioxapentadecane, 2,5,8,11-tetraoxadodecane, and 2,5,8,11,14-pentaoxapentadecane are reported from densities measured with a vibrating-tube densimeter. Systems containing 2,5-dioxahexane, 2,5,8-trioxanonane, 2,5,8,11-tetraoxadodecane or 2,5,8,11,14-pentaoxapentadecane are characterized by VmE > 0, probably due to predominant positive contributions to VmE from the disruption of H bonds of 1-hexanol and to physical interactions. In contrast, mixtures with 3,6-dioxaoctane, 3,6,9-trioxaundecane, and 5,8,11-trioxapentadecane are characterized by VmE < 0, indicating that the negative contribution to VmE from interstitial accommodation is more important.  相似文献   

20.
13 Spectra have been studied and signals of C atoms assigned for steroids 3-5 and 7-10, which are close structural analogs of the natural ecdysteroids 4-dehydroecdysterone 1, diaulusterol A 2, and 5,20-dihydroxyecdysone 6  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