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.
《Fluid Phase Equilibria》1987,32(2):205-210
The solvation behaviour of silver(I) iodate in methanol—acetonitrile (AN) and ethanol—acetonitrile mixtures has been studied at 30°C by solubility and emf measurements. The solubility of the salt increases with the addition of AN and passes through a maximum at XAN = 0.3 and 0.6 in the case of MeOH-AN and EtOH-AN mixtures, respectively, and then decreases with further addition of AN. The transfer free energy of silver ion decreases while that of iodate ion increases with the addition of AN in both the solvent mixtures. The solvent transport number, Δ of AN is positive with a maximum at XAN = 0.45 (Δ = 0.45 (Δ = 5.4) and at XAN = 0.55 (Δ = 2.4) in the case of MeOH-AN and EtOH-AN mixtures, respectively. These results have been interpreted in terms of the heteroselective solvation of the salt, the silver ion being preferentially solvated by AN and the iodate ion by the amphiprotic solvent component in these mixtures.  相似文献   

2.
The selective solvation of copper(I) thiocyanate has been studied in water-acetonitrile (AN) mixtures at 30°C by solubility and EMF measurements. The solubility of the salt increases continuously and changes only slightly beyond X AN =0.7 mole fraction. The Gibbs energy of transfer of copper(I) ion from water to mixtures with acetonitrile (determined on the basis of the ferrocene reference method) decreases continuously, while that of the thiocyanate ion increases with the addition of acetonitrile. The solvent transport number of acetonitrile passes through a maximum (=6.4) at X AN =0.25. These results were interpreted as arising due to a heteroselective solvation of the salt, the copper(I) ion being preferentially solvated by acetonitrile and the thiocyanate ion by water in these mixtures.  相似文献   

3.
The solvent transport number, , of dimethyl sulfoxide andGibbs solvation energies of silver acetate in the binary solvent systems, water—DMSO and methanol—DMSO, were determined by employingEMF and solubility measurements. While the transfer free energy of the salt increases from water to water—DMSO mixtures (up toX DMSO =0.7) and then decreases, it continuously decreases from methanol to methanol—DMSO mixtures. In both mixed solvents, t(Ag+) decreases down to pureDMSO and that of acetate ion increases with increasing composition ofDMSO indicating that silver ion is preferentially solvated byDMSO and acetate ion by water or methanol in these mixtures. The solvent transport numbers, , ofDMSO are positive throughout, passing through a maximum atX DMSO =0.45 (=1.0) in the case of water—DMSO mixtures and atX DMSO =0.25 (=1.8) in methanol—DMSO mixtures. This observation is shown to be in accord with the conclusions arrived at from the transfer energy data of the salt in the two mixtures.
Bevorzugte Solvatation von Silber(I) acetat in Wasser, Methanol und deren Mischungen mit Dimethylsulfoxid
Zusammenfassung Es wurde die Lösungsmitteltransportzahl () von Dimethylsulfoxid und dieGibbssche Solvatationsenergie von Silberacetat in den binären Lösungsmittelsystemen Wasser—DMSO und Methanol—DMSO mittelsEMK- und Löslichkeitsmessungen ermittelt. Während die freie Energie des Transfers des Salzes beim Übergang von Wasser zu Wasser—DMSO-Mischungen zunächst ansteigt (bisX DMSO =0,7) und dann abfällt, ist für das System Methanol bzw. Methanol—DMSO ein kontinuierlicher Anstieg feststellbar. In beiden Mischsystemen fällt G°t(Ag+) bis zu reinemDMSO, der entsprechende Wert für Acetat steigt bei großeren Anteilen vonDMSO an. Das zeigt, daß das Silber(I)ion bevorzugt vonDMSO solvatisiert wird, Acetat von der jeweils anderen Komponente (Wasser bzw. Methanol). Die Lösungsmitteltransportzahl fürDMSO ist stets positiv mit Maxima beiX DMSO =0,45 (=1,0) für Wasser—DMSO undX DMSO =0,25 (=1,8) für Methanol—DMSO. Diese Beobachtung steht im Einklang mit Rückschlüssen, die aus den Transferenergiedaten des Salzes in den zwei untersuchten Systemen getroffen werden können.
  相似文献   

4.
The stability constants of iron(III) complexes with nicotinamide in water-DMSO mixtures (X DMSO = 0–0.75) were determined by potentiometric titration at 25.0 ± 0.1°C and an ionic strength of 0.25 (NaClO4). The contributions from the solvation of the reagents to the Gibbs energy of complexation transfer were analyzed. The stabilities of iron(III), copper(II), and silver(I) complexes with nicotinamide were compared. The observed decrease in the stability constants was attributed to the stabilization of iron(III) solvate complexes as the DMSO content increases.  相似文献   

5.
The effect of a water–dimethyl sulfoxide solvent (X DMSO= 0–0.97, where X DMSOis the mole fraction of DMSO) on the thermodynamics of complexation between Ag+and 18-crown-6 and the solvation of all reagents involved in this equilibrium were studied. In aqueous solutions, the complex is stable mainly because of the enthalpy contribution to r G°. For X DMSO> 0.3, the contributions from entropy and enthalpy become comparable in magnitude, but they are opposite in sign. In the binary solvent, the complex is most stable at X DMSO= 0.2 to 0.3. Analysis of the enthalpy characteristics of reagent solvation showed that this solvent effect was due to the superposition of two opposite solvation contributions occurring with an increase in the DMSO concentration in the binary solvent, namely, the destabilization of the ligand solvate sphere and the formation of stable Ag+complexes with DMSO.  相似文献   

6.
The effect of composition of ethanol–dimethyl sulfoxide (EtOH–DMSO) solvents (χDMSO = 0.0–1.0 mole fractions) on the stability of silver(I) complexes with 18-crown-6 ether (18C6) has been studied potentiometrically at 298.15 K. The increasing of DMSO concentrations in mixed solvents are shown to considerably reduce the stability of 18C6 complexes with silver(I) ion ([Ag18C6]+). A change in the solvation state of the central ion is suggested to be the key factor in shifting complexing equilibrium.  相似文献   

7.
The ionization enthalpy of benzoic acid has been measured calorimetrically at 25°C in H2ODMSO mixtures ranging from pure water to a maximum DMSO molar ratio XDMSO = 0.80. With the increase of DMSO content, the ionization becomes more and more endothermic, and for XDMSO = 0.8 the ionization enthalpy is about 6 kcal mol?1 higher than in water. By also measuring the solution enthalpy of crystalline benzoic acid in the mixtures, it has been shown that the solvation of the undissociated molecule is the main cause for the increase of the dissociation enthalpy. A comparison has been made between the relative enthalpies of benzoic and hydroxide ions in H2ODMSO mixtures.  相似文献   

8.
The solvation state of biologically active compound vitamin B3, viz., 3-pyridinecarboxamide, in an aqueous-dimethyl sulfoxide solvent of a variable composition was studied by 1H and 13C NMR and IR spectroscopy. Below X DMSO ?0.65 molar fraction, the solvation of the N heteroatom due to hydrogen bonds with water molecules weakens. At X DMSO > 0.65 molar fraction, almost no changes are observed in the solvate state of the N heteroatom. The 1H NMR spectra indicate that the degree of conjugation of the carbamide group with the heterocycle increases with an increase in the DMSO concentration. The structures of the dimethyl sulfoxide and mixed aqueous-dimethyl sulfoxide solvates of nicotinamide were optimized by the B3LYP/6311++(DP) method, and their 13C chemical shifts (GIAO) and IR spectra were obtained. According to the IR spectroscopic data, the number of hydrogen bonds involving the carbamide group decreases on going from H2O to DMSO.  相似文献   

9.
The Gibbs energies of transfer of 18-crown-6 ether from water into water-dimethyl sulfoxide (DMSO) solvents (χDMSO = 0.0–0.97 mole fractions) at 298.15 K were determined by the interphase distribution method. Changes in the composition of the aqueous-organic solvent did not cause noticeable changes in the stability of 18-crown-6 ether solvato complexes. Reagent solvation contributions to shifts of complex formation equilibrium between silver(I) and 18-crown-6 ether when water was replaced with dimethyl sulfoxide were analyzed.  相似文献   

10.
The stability change of nickel(II) ion complexes including one and two nicotinamide (B3 vitamin) molecules in aqueous dimethyl sulfoxide (XDMSO = 0–0.85 m.f.) was studied at 298.2±0.1 K and 0.25 ionic strength value (NaClO4) using the potentiometric method. The first stage constant of complexation increased until organic solvent concentration was 0.5 m.f. and reduced at higher DMSO content. The difference between complex and central ions solvation is a dominating contribution into the Gibbs energy change of mononicotinamide complex formation reaction. When the second ligand molecule was bonded into the coordination compound, the nicotinamide contribution to ΔtrGr rose and became prevailing at XDMSO = 0.7–0.85. The ligand was found to replace a water molecule in the coordination sphere of the cation according to spectrophotometric study results.   相似文献   

11.
The Gibbs energies of transferring triglycine (3Gly, glycyl-glycyl-glycine) from water into mixtures of water with dimethyl sulfoxide (χDMSO = 0.05, 0.10, and 0.15 mole fractions) at 298.15 K are determined from the interphase distribution. An increased dimethyl sulfoxide (DMSO) concentration in the solvent slightly raises the positive values of Δtr G (3Gly), possibly indicating the formation of more stable 3Gly-H2O solvated complexes than ones of 3Gly-DMSO. It is shown that the change in the Gibbs energy of transfer of 3Gly is determined by the enthalpy component. The relationship of 3Gly and 18-crown-6 ether (18C6) solvation’s contributions to the change in the Gibbs energy of [3Gly18C6] molecular complex formation in H2O-DMSO solvents is analyzed, and the key role of 3Gly solvation’s contribution to the change in the stability of [3Gly18C6] upon moving from H2O to mixtures with DMSO is revealed.  相似文献   

12.
The standard Gibbs transfer energies of the silver(I)-18-crown-6 perchlorate complex salt from methanol to various compositions of methanol-acetonitrile mixtures were determined from solubility measurements at 30°C and these data were separated into the corresponding ionic contributions by employing the negligible liquid junction potential method of Parkeret al. The solvent transport numbers AN, for the salt were also determined at various solvent compositions using a concentration cell with transference.The Gibbs transfer energy of the silver(I)-18-crown-6 complex cation is negative and decreases with the addition of acetonitrile but the transfer energy of the anion is positive and increases under the same conditions. The solvent transport number, AN, increases and passes through a maximum value of 5.48 at AN=0.55. These results indicate that the complex salt is heteroselectively solvated in these mixtures with the cation being preferentially solvated by acetonitrile and the anion by methanol molecules.  相似文献   

13.
The equilibrium solubility and preferential solvation of triclocarban in {1,4-dioxane (1) + water (2)} mixtures at 298.15 K was reported. Mole fraction solubility varies continuously from 2.85 × 10–9 in neat water to 2.39 × 10–3 in neat 1,4-dioxane. Solubility behaviour was adequately correlated by means of the Jouyban-Acree model. Based on the inverse Kirkwood-Buff integrals, preferential solvation parameters were calculated. Triclocarban is preferentially solvated by water in water-rich mixtures (0.00 < x1 < 0.18) and also in 1,4-dioxane-rich mixtures (0.78 < x1 < 1.00) but preferentially solvated by 1,4-dioxane in mixtures with similar solvent compositions.  相似文献   

14.
A method is described for the estimation of iodide, based on its oxidation to iodate by the addition of excess of chloramine-T, destruction of the excess of chloraniine-T with dimethyl sulphoxide and determination of the iodate iodometrically. In addition, the dimethyl sulphoxide eliminates any interference from bromide or bromate.  相似文献   

15.
The solvation dynamics of three coumarin dyes with widely varying polarities were studied in acetonitrile–water (ACN–H2O) mixtures across the entire composition range. At low ACN concentrations [ACN mole fractions (XACN)≤0.1], the solvation dynamics are fast (<40 ps), indicating a nearly homogeneous environment. This fast region is followed by a sudden retardation of the average solvation time (230–1120 ps) at higher ACN concentrations (XACN≈0.2), thus indicating the onset of nonideality within the mixture that continues until XACN≈0.8. This nonideality regime (XACN≈0.2–0.8) comprises of multiple dye‐dependent anomalous regions. At very high ACN concentrations (XACN≈0.8–1), the ACN–H2O mixtures regain homogeneity, with faster solvation times. The source of the inherent nonideality of the ACN–H2O mixtures is a subject of debate. However, a careful examination of the widths of time‐resolved emission spectra shows that the origin of the slow dynamics may be due to the diffusion of polar solvent molecules into the first solvation shell of the excited coumarin dipole.  相似文献   

16.
The stability of complexes and enthalpy of interaction of Ag+ ions with 18-crown-6 in waterdimethyl sulfoxide (DMSO) mixtures were determined by calorimetric titration in the range of mole fractions XDMSO from 0.0 to 0.97 at 298.15 K. With increasing concentration of the nonaqueous component in the solvent to XDMSO 0.3, the stability of the complex ion [AgL]+ increases, which is followed by a decrease in logK(AgL+) to 0.35 plusmn 0.15 at XDMSO 0.97. The exothermic effect of the reaction shows a similar trend. The presence of the extremum in the logK-XDMSO and r H-XDMSO dependences is explained by the competition of two solvation contributions: destabilization of the ligand with decreasing water content in the solvent and formation of strong solvation complexes of Ag+ with DMSO.  相似文献   

17.
Stability constants of nickel(II) glycylglycinate complexes in aqueous solutions of dimethylsulfoxide of variable composition (from 0.00 to 0.60 mole fractions DMSO) are determined according to potentiometry at 298.15 K and an ionic strength of 0.1 M (NaClO4). It is determined that with a rise in the concentration of an organic cosolvent in solution, the stability of nickel(II) complexes with glycylglycinate ion on the whole increases, but the logK stability = f(X DMSO) dependences are of a critical character with a maximum of 0.3 mole fractions DMSO. It is demonstrated that the rise in the stability of complexes is related to the destabilization of ligands in the low concentration range of the organic component, while the presence of a maximum is due to the different dynamics of the solvation contributions from reagents during changes in the Gibbs energy of reaction.  相似文献   

18.
Summary Compounds of the type RuL4–nX2+n where L = dimethyl sulphoxide (DMSO), tetramethylene sulphoxide (TMSO) and X = Cl, Br or I for n = 0 and L = di-n-propyl sulphoxide (n-Pr2SO) and di-n-butyl sulphoxide (n-Bu2SO) X = Cl or Br for n = 1 and also Ru(DMSO)6Br3 have been prepared and studied. The important i.r. bands of all compounds together with their electronic spectra and the thermograms of some of them are discussed. In order to interpret the i.r. data, the corresponding deuteriated (DMSO-d6) analogues have also been prepared. In the majority of the compounds of the type RuL4X2 the sulphoxide ligands are bonded through the sulphur atom; in a few cases, bonding through both S- and O-donor sites has been found. A mixed type of bonding is also observed in Ru(DMSO)6Br3 and in RuL3X3.  相似文献   

19.
The mole fraction solubility of phenacetin (PNC) in methanol + water binary solvent mixtures at 298.15 K was determined along with density of the saturated solutions. All these solubility values were correlated with the Jouyban–Acree model. Preferential solvation parameters of PNC by methanol (δx1,3) were derived from their thermodynamic solution properties using the inverse Kirkwood–Buff integrals (IKBI) method. δx1,3 values are negative in water-rich mixtures but positive in methanol mole fraction of >0.32. It is conjecturable that in the former case the hydrophobic hydration around non-polar groups of PNC plays a relevant role in the solvation. The higher solvation by methanol in mixtures of similar cosolvent compositions and methanol-rich mixtures could be explained in terms of the higher basic behaviour of methanol.  相似文献   

20.
The stability constants, β1, of each monochloride complex of Ln(III) (Ln=Nd or Tm) have been determined in the mixed system of dimethyl sulfoxide (DMSO) and water with 1.0 mol·dm−3 ionic strength using a solvent extraction technique. The values of β1 of Ln(III) decrease to about 0.2 mole fraction of DMSO (X s) in the mixed solvent system and then increase withX s (>about 0.2). However, the variation mode of β1 of Nd(III) withX s somewhat differs from that of Tm(III). Calculation of Ln3+−Cl distance using a Born-type equation of the Gibbs' free energy derived from the β1 evealed the followings: (1) For Tm3+ with coordination number 8, the estimated distance between Tm3+ and Cl (d Tm-Cl) increases linearly withX s in 0.00≤X s≤0.17. This means an enlargement of the primary solvation sphere size of Tm3+ withX s. On the other hand, thed Tm-Cl shows a decrease withX s in 0.17<X s<0.28. (2) The estimatedd Nd-Cl increases linearly withX s in 0.00≤X s<0.06 and 0.06<X s≤0.17, but their slopes are different. The larger slope againstX s in 0.06<X s≤0.17 is attributable to a lowering of the β1 by a coordination of ClO4 into the secondary solvation sphere of Nd3+ and/or by an increase in the solvation number of the primary solvation sphere of Nd3+.  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