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 546 毫秒
1.
The solubility of trans-(Coen2Cl2)2ReCl6 has been determined in water and in water +t-butyl alcohol mixtures. By comparzng these values with the solubility of Cs2ReCl6 in similar mixtures, values for the difference in free energy of transfer, G t o (i) between water and water + t-butyl alcohol can be calculated for i =[Coen2Cl2]+ and Cs+. The introduction of G t o (Cs+) then produces values for G t o (Coen2Cl2 +). The difference in G t o (i) for i=[Coon2Cl]2+ in the transition state and i=[Coen2Cl2]+ in the initiad' tate for the solvolysis of the trans-[Coen2Cl2]+ ion in water + t-butyl alcohol can be derived from the application of a free energy cycle: using G t o (Coen2Cl2 +) determined from the solubility measurements allows the calculation of values for G t o (Coen2Cl2+). G t o (i) in water + t-butyl alcohol for bis (1,2-diamino) cobalt (III) ions are compared with G t o (i) for tetrapyridinecobalt (III) ions.  相似文献   

2.
The solubility of salts [Co(3Rpy)4Cl2]2]ReCl6] has been determined in water + methanol mixtures. By comparing these with the solubilities of the salt Cs2ReCl6 and using calculated activity coefficients for the ions in the water+methanol mixtures, values for {G t o (Co(3Rpy)4Cl 2 + )–G t o (Cs+)} can be determined where G t o is the standard Gibbs free energy of transfer from water to an aqueous mixture. G t o (Cs+) from the solvent sorting scale and from the TPTB scale are then used to calculate G t o (Co(3Rpy)4Cl 2 + ). These two sets of values for G t o (Co(3Rpy)4Cl 2 + ) on the differing scales are then inserted into a free energy cycle applied to the bond extension Co(3Rpy)4Cl 2 + (initial state)Co(3Rpy)4Cl2++Cl (transition state) for the solvolysis in water and in water + methanol mixtures to produce values for G t o (Co(3Rpy)4Cl2+) using both scales. Data for the solubilites of [Copy4Cl2]2[ReCl6] and [Co(4Rpy)4Cl2]2[ReCl6] have been re-calculated to compare free energies of transfer for these complex cations with those specified above.  相似文献   

3.
The kinetics of the solvolysis of Co(CN)5Cl3– have been investigated in water with an added structure former, ethanol, and with added urea, which has only a weak effect on the solvent structure. As this solvolysis involves a rate-determining dissociative step corresponding closely to a 100%; separation, Co3+ Cl-, in the transition state, a Gibbs energy cycle relating Gibbs energies of activation in water and in the mixtures to Gibbs energies of transfer of individual ionic species between water and the mixtures, G t o (i), can be applied. The acceleration of the reaction found with both these cosolvents results from the compensation of the retarding positive G t o (Cl- by the negative term [G t o [Co(CN) 5 2- ]-G t o [Co(CN)5Cl3- arising from G t o [Co(CN)5Cl3-]> G t o [Co(CN) 5 2- ]. Moreover, only a small tendency to extrema in the enthalpies and entropies of activation is found with both these cosolvents, as was also found with added methanol or ethane-1,2-diol, but unlike the extrema found when hydrophobic alcohols are added to water. With the latter, much greater negative values for G t o [Co(CN) 5 2- ]- t o [Co(CN)5Cl3-] are found. When G G t o [Co(CN) 5 2- ]-G G t o [CO(CN)5Cl3-] becomes low enough not to compensate for the positive G t o (Cl-), as with added hydrophilic glucose, the reaction is retarded. Compensating contributions of the various G t o (i) involved in the Gibbs energy cycle with added methanol or ethane-1, 2-diol allow log (rate constant) to vary linearly with the reciprocal of the relative permittivity of the medium.  相似文献   

4.
A recently introduced modified hydration shell hydrogen bond model for rationalizing the thermodynamic consequences of hydrophobic hydration is adapted for use with heavy water. The required adjustment of parameters employs the assumption that breaking hydrogen bonds in water-d2 involves a greater enthalpy change and a larger entropy increase than bond breaking in ordinary water. It also makes some use of information derived from studies of gas solubilities in the two solvents, although a review of the data leads to serious questions about the reliability of results obtained in this way. The model permits calculations of hydrogen bonding contributions to the changes, G t o , H t o , S t o , and C p,t o , for transfer of nonpolar solutes from water to water-d2 and implies that such data should show regular trends. Although some of the numerical results depend strongly on the values chosen for the parameters, the pattern defined by these trends is nearly independent of parameters. Predicted values of C p,t o are large and positive for all nonpolar solutes, while S t o is expected to be negative near 0°C, becoming progressively less negative on warming and eventually positive. Both of these quantities should be proportional to the molecular surface area of the solute. Analogous predictions regarding G t o and H t o can also be made, but only if it is permissible to neglect possible contributions to these quantities from van der Waals interactions.  相似文献   

5.
Heats of solution of nine electrolytes in 1,2-dichloroethane and of three electrolytes in 1,1-dichloroethane have been determined calorimetrically at various electrolyte concentrations and extrapolated to zero concentration to yield H s o values for these electrolytes. It is shown that values of H t o for transfer from water to the dichloroethanes of 11 electrolytes are often negative, so that these electrolytes can be more stable enthalpically in the less polar solvents. Combinations of the H t o values with previously determined G t o values yield values of S t o for transfer of 11 electrolytes from water to the dichloroethanes. These S t o values are mostly very negative; they can be correlated very well by the method of Abraham, and in this way S t o values for transfer of numerous other anions and cations have been predicted. The Ph4As+/Ph4B convention yields single-ion entropies of transfer from water to the dichloroethanes in reasonable agreement with values calculated by the correspondence-plot method.  相似文献   

6.
Rates of solvolysis of the complex cation [Co(4tBupy)4Cl2]+ have been determined in mixtures of water with the hydrophobic solvent, t-butyl alcohol. The solvent composition at which the extremum is found in the variation of the enthalpy H* and the entropy S* of activation correlates well with the extremum in the variation of the relative partial molar volume of t-butyl alcohol in the mixture and the straight line found for the variation of H* with S* is coincident with the same plot for water + 2-propanol mixtures. A free energy cycle is applied to the process initial state (C n+) going to the transition state [M(n+1)+...Cl] in water and in the mixture using free energies of transfer of the individual ionic species, G t o (i), from water into the mixture. Values for G t o (i) are derived from the solvent sorting method and from the TATB/TPTB method: using data from either method, changes in solvent structure on going from water into the mixture are found to stabilize the cation in the transition state, M(n+1)+, more than in the initial state, C n+. This is compared with the application of the free energy cycle to the solvolysis of complexes [Co(Rpy)4Cl2]+ and [Coen2LCl]+ in mixtures of water with methanol, 2-propanol or t-butyl alcohol: the above conclusion regarding the relative stabilization of the cations holds for all these complexes in their solvolyses in water+alcohol mixtures using values of G t o (Cl) from either source.  相似文献   

7.
Summary The solvolysis oftrans-[Co(4-Etpy)4Cl2]ClO4, was followed spectrophotometrically in water/isopropanol at different temperatures. The activation energy varied nonlinearly with the mole fraction of the co-solvent, 2. The plot of logk versus D s –1 was also non-linear. These features were attributed to the differential solvation of the initial and transition states. On plotting H versus S, the points fall very close to straight line. The isokinetic temperature was found to be 334K, indicating that the solvolysis reaction is controlled by S and not H. The change in H and S with the mole fraction of the cosolvent shows extrema at the composition range where changes in solvent structure occur. The influence of the solvent structure on the complex ion in the transition state dominates over that in the initial state, where –G t 0 [Co(4-Etpy)4Cl]2+>–G t 0 [Co(4-Etpy)4Cl2]+.  相似文献   

8.
A recently developed calorimetric method has been employed to estimate the thermodynamic functions for transfer of 1-propanol, 1-butanol, 1-pentanol, 1-hexanol and 1-heptanol from n-octane to water at 25°C. A linear correlation for G t o as a function of the number of carbon atoms of the alchohol molecule has been found but for H t o and S t o the dependence gave well defined minima.  相似文献   

9.
The standard potentialss E o of M/M+ (M=Li, Na, K, Rb, and Cs) electrodes in aqueous urea solutions containing 12, 20, 30 and 37% by weight of urea have been determined at 25°C from emf measurements on the cell M(Hg)/MCl (m), solvent/AgCl–Ag, from the activities of metals in amalgams by use of a similar type of cell in water, and from the values ofs E o of the Ag/AgCl electrode determined earlier. The standard free energies of transfer of MCl, G t o (MCl), from water to the mixed solvents, computed by use of these values and those for the Ag–AgCl electrode, rise sharply from Li+ to Na+ but fall from Na+ to K+ and rather sharply from K+ to Cs+ with a maximum at Na+ in all the solvent compositions. This has been attributed to the superimposition of soft-soft interactions on the electrostatic interactions between the ions and the negative charge centers of the possible hydrogen-bonded solvent complexes in the mixed solvents. Comparison of G t o (i) values for individual ions, obtained by a simultaneous extrapolation procedure, with those in aqueous mixtures of methanol,t-butanol, and dimethyl sulfoxide leads to the conclusion that the solvation of these ions in all these solvents is chiefly dictated by the acid-base type of ion-solvent interactions.  相似文献   

10.
NMR and hydrogen equilibrium pressure measurements were performed on hydrides of the intermetallic compounds Ti2(Ni, Co) and Ti2(Ni, Fe). The following values of enthalpy H and entropy S for the formation of the hydrides of the intermetallic phases Ti2Co and Ti2Ni were found: H(Ti2CoH y )=–47.6 kJ/mol H2, H(Ti2NiH y )=–53.7 kJ/mol H2; S(Ti2CoH y )=–119.8 J/(K·mol H2), S(Ti2NiH y )=–127.5 J/(K·mol H2). By substitution of Ni or Co by Fe, the values of H and S of the corresponding quaternary hydrides become less negative. An interpretation of the experimental results is tried by the model ofShaltiel and coworkers.Proton diffusion was investigated in a series of the intermetallic hydrides Ti2(Ni, Co)H x and Ti2(Ni, Fe)H x . The diffusion rate is lowered by increased Ni/Fe substitution. Substitution of Ni by Co scarcely effects the hopping process. The activation energies were found to be smaller for the Ti2Ni-hydrides compared with the Ti2Co-hydrides.
Herrn Prof. Dr.H. Nowotny zum 70. Geburtstag gewidmet.  相似文献   

11.
Integral enthalpies of solution of several amino acids in water at low concentrations have been determined at 25 and 35°C. These data have been used to derive the heat-capacity change C p o on dissolution at 30°C. Partial molal heat capacities C p2 o have been obtained by combining C p o with C p2 o (heat capacity of pure solid amino acids). The results indicate that the increments in C p o and C p2 o values per CH2 group increment in the homologous series of -amino acids are constant and in agreement with those found for other homologous series of compounds containing monofunctional groups. However, this is not the case with amino acids having the NH 3 + group at the terminal position. The present work also indicates that, as the NH 3 + group is shifted away from the COO group, hydrophobic hydration decreases, as indicated by a decrease in C p o and C p2 o . the results on various isomers of amino acids show that branching of alkyl groups has no effect on C p o and C p2 o , indicating that hydrophobic hydration is unaffected by branching. The effect of substitution of H by OH and of CH3 by groups in some amino acids has also been studied and discussed.  相似文献   

12.
Summary The stoichiometric stability constants for La(III) and Y(III)L-serine complexes were determined by potentiometric methods at different ionic strengths adjusted with NaClO4 and at different temperatures. The overall changes in free energy (G o), enthalpy (H o), and entropy (S o) during the protonation ofL-serine and that accompanying the complex formation with the metal ions have been evaluated.
Komplexbildungskonstanten und thermodynamische Parameter für La(III)- und Y(III)-L-Serin-Komplexe
Zusammenfassung Die stöchiometrischen Komplexbildungskonstanten für La(III)- und Y(III)-L-Serin-Komplexe wurden mittels potentiometrischer Methoden bei verschiedenen Ionenstärken (mit NaClO4 adjustiert) und bei verschiedenen Temperaturen bestimmt. Die Änderungen in der freien Energie (G o), Enthalpie (H o) und Entropie (S o) während der Protonierung und der Komplexbildung mit den Metallionen wurden ermittelt.
  相似文献   

13.
The infinite-dilution apparent molar volumesV 2 o for glycine, DL-alanine, DL--amino-n-butyric acid, DL-valine, DL-leucine, and L-serine in 6 mol-kg–1 aqueous guanidine hydrochloride were determined at 5, 15, 25, and 35°C from precise density measurements. Using these data, the standard volumes of transfer, t , from water to 6m> aqueous guanidine hydrochloride solution were calculated. A linear relationship was found between V 2 o and temperature. Both V 2 o and t vary linearly with increasing number of carbon atoms in the alkyl chain of the amino acids. The results show that the apparent molar volumes at infinite dilution for (NH 3 + ,COO-) groups increase with increasing temperature and those for CH2 and the other alkyl chains are almost constant. These results also shows that guanidine hydrochloride has stronger interactions with amino acids than urea. These phenomena are discussed in terms of the cosphere overlap model.  相似文献   

14.
Heats of solution of 13 11 electrolytes in 1-propanol have been determined calorimetrically at various electrolyte concentrations, and extrapolated to zero concentration to give H s o values for these electrolytes. Together with literature data on three additional 11 electrolytes, these measurements yield a self-consistent set of single-ion enthalpies of transfer from water to 1-propanol. Values are tabulated for 10 univalent cations and five univalent anions. It is shown that the H t o (Ph 4 As+)=H t o (Ph 4 B) assumption yields chemically reasonable single-ion values. Using this assumption, it may be deduced that all the univalent ions studied have about the same enthalpy in 1-propanol as in methanol.  相似文献   

15.
Enthalpies of solution of thymine and uracil in water and in dimethylsulfoxide (DMSO) were measured calorimetrically in the temperature range 25–40°C. H s o at 25°C for thymine and uracil in water were found to be 23.1±0.5 and 29.5±0.3 kJ-mol–1, respectively. In DMSO, H s o were 7.9±0.1 and 10.2±0.1 kJ-mol–1, respectively. In aqueous solution C p o for the two nucleic acid bases were relatively large and positive with C p o of thymine being larger. Both transfer quantities H t o and C p,t o for the proceses H2ODMSO for the two nucleic acid bases were negative. It is proposed that, the differences in the values obtained for the two bases is due principally to increased order in the water adjacent to the methyl group in thymine.  相似文献   

16.
The interaction of thymidine, a nucleoside, with hydroxopentaaquarhodium(III), [Rh(H2O)5(OH)]2+ ion in aqueous medium is reported and the possible mode of binding is discussed. The kinetics of interaction between thymidine and [Rh(H2O)5OH]2+ has been studied spectrophotometrically as a function of [Rh(H2O)5OH2+], [thymidine], pH and temperature. The reaction has been monitored at 298 nm, the max of the substituted complex, and where the spectral difference between the reactant and product is a maximum. The reaction rate increases with [thymidine] and reaches a limiting value at a higher ligand concentration. From the experimental findings an associative interchange mechanism for the substitution process is suggested. The activation parameters (H=47.8 ± 5.7 kJ mol–1, S=–173 ± 17 J K–1 mol–1) supports our proposition. The negative G0 (–13.8 kJ mol–1) for the first equilibrium step also supports the spontaneous formation of the outer sphere association complex.  相似文献   

17.
The kinetics of the solvolysis of [Co(CN)5Cl]3– have been investigated in water +2-methoxyethanol and water + diethylene glycol mixtures. Although the addition of these linear hydrophilic cosolvent molecules to water produces curvature in the variation of log(rate constant) with the reciprocal of the dielectric constant, their effect on the enthalpy and entropy of activation is minimal, unlike the effect of hydrophobic cosolvents. The application of a Gibbs energy cycle to the solvolysis in water and in the mixtures using either solvent-sorting or TATB values for the Gibbs energy of transfer of the chloride ion between water and the mixture shows that the relative stability of the emergent solvated Co(III) ion in the transition state compared to that of Co(CN)5Cl3– in the initial state increases with increasing content of cosolvent in the mixture. By comparing the effects of other cosolvents on the solvolysis, this differential increase in the relative stabilities of the two species increases with the degree of hydrophobicity of the cosolvent.List of Symbols v2 partial molar volume of the cosolvent in water + cosolvent mixtures - V 2 o molar volume of the pure cosolvent - H mix E excess enthalpy of mixing water and cosolvent - S mix E excess entropy of mixing water and cosolvent - G t o (i)n the Gibbs energy of transfer of speciesi from water into the water + cosolvent mixture excluding electrostatic contributions - k s first order rate constant for the solvolysis in water + cosolvent mixtures - D s dielectric constant of the water + cosolvent mixture - H * the enthalpy of activation for the solvolysis - S * the entropy of activation for the solvolysis - G * the Gibbs energy of activation for the solvolysis - V * the volume of activation for the solvolysis - i * speciesi in the transition state for the solvolysis - H o Hammett Acidity Function - TATB method for estimating the Gibbs energy of transfer for single ions assuming those for Ph4As+ and BPh 4 are equal  相似文献   

18.
Variations of densities and viscosities with temperature and composition are reported for binary liquid mixtures containing propionic acid+aniline (I),+o-toluidine (II),+o-anisidine (III), and+o-chloroaniline (IV). Entropies S m and enthalpies H m of activation as functions of the composition of the mixtures, as well as free energies of activation G m at 10, 20, 30, 40, and 50°C and different compositions were calculated by means of Eyring's equation. The formation of activated complexes between the components of these binary mixtures is postulated and claimed to result from acid-base and hydrogen bonding exchange interactions.  相似文献   

19.
Conclusions The relationship between the free energies of activation G and reaction Go for proton transfer processes have been analyzed, taking into account the effect of hindered rotation of the reagents. We have shown that the considered effect can considerably affect the shape of the G=f(Go) curve.Translated from Izvestiya Akademii Nauk SSSR, Seriya Khimicheskaya, No. 1, pp. 77–81, January 1989.  相似文献   

20.
Summary The kinetics of the thermal and photochemical decomposition of aquapentacyanoferrate(III) ion in aqueous solution in the presence ofo-phenanthroline was studied spectrophotometrically. The first-order rate constant (k ) at 30° C [I=1 M(NaCl)] for the thermal reaction is (1.49±0.13)×10–6 s–1 with H =(158±7)kJ mol–1 and S=(42±4) JK–1 mol–1. The initial quantum yield for the photochemical reaction at pH=7 is independent of the light intensity and is (1.49±0.33)×10–2 mol einstein–1.A communication on this subject was presented at the XVI Latinamerican Chemistry Congress held at Rio de Janeiro. Brasil, October 14–20, 1984.  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