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 31 毫秒
1.
A series of concentrated aqueous solutions of ferric chloride with different chloride:iron(III) ratios has been studied by means of EXAFS to determine the structure around the iron(III) ion of the dominating species in such solutions. The dominating species in dilute acidic aqueous solution of ferric chloride, at less than 1 mmol·dm?3, are the hydrated iron(III) and chloride ions, while in concentrated aqueous solution and in solutions with an excess of chloride ions, up to 1.0 mol·dm?3, it is the trans-[FeCl2(H2O)4]+ complex. Possible higher chloroferrate(III) or dimeric [Fe2Cl6] complexes at room temperature, as proposed in the literature, were not observed in any of the studied solutions in spite of an excess of chloride ions of 1 mol·dm?3.  相似文献   

2.
Nernst—Planck equations and ionic conductivities are used to calculate accurate limiting interdiffusion coefficients D ik o for mixed electrolyte solutions. The electrostatic mechanism for coupled electrolyte diffusion is investigated by calculating the electrostatic contribution to each D ik o coefficient to give the flux of each electrolyte driven by the electric field, which is generated by the migration of ions of different mobilities. Ternary diffusion coefficients are measured for dilute aqueous K2SO4 + KOH and Li2SO4 + LiOH solutions. Because of the different mobilities of K+ and Li+ ions relative to SO 4 2– ions, diffusing K2SO4 drives cocurrent flows of KOH, but diffusing Li2SO4 drives counterflows of LiOH. To describe coupled diffusion in concentrated mixed electrolyte solutions, the Hartley–Crank theory is used to correct the limiting D ik o coefficients for nonideal solution behavior, viscosity changes, ionic hydration, and the zero-volume flow constraint. Diffusion coefficients predicted for concentrated aqueous CaCl2 + HCl solutions are compared with recently reported data. The large amount of HCl cotransported by the diffusing CaCl2 is attributed to the salting out of HCl by CaCl2 and to the migration of H+ ions in the diffusion-induced electric field, which slows down the Cl ions and speeds up the less-mobile Ca2+ ions to maintain electroneutrality along the CaCl2 gradient.  相似文献   

3.
Intrinsic ionic heats of transport q o * (ion) and ionic heats of transport Q o * (ion) have been evaluated for 53 aqueous ions at infinite dilution at 25°C using the reduction rule proposed by the authors and the limiting laws of Agar, and of Helfand and Kirkwood without electrophoretic terms. q o * (ion) have been found to correlate linearly with the standard ionic entropies of hydration for the 38 ions investigated. The correlation yields three distinctive proportionality constants indicating that the ions may be divided into three distinctive groups. Although the sign of Q o * (ion) is not definite, all values of q o * (ion) are positive. For 17 ions Q o * (ion) are in good agreement with TS o * (ion). Here, S o * (ion) is the absolute standard ionic entropy of transport which can be obtained from potentiometric measurements on cells. The values of S o * (ion) were determined by Agar, and recently by Lin and coworkers.  相似文献   

4.
Molar conductances of dilute aqueous benzoic acid solutions are presented for temperatures from 5 to 80°C. The data have been analyzed to give acid dissociation constants as well as ΔH o, ΔS o, and ΔC p o for the ionization process and the limiting conductance of the benzoate ion. The conductance-viscosity product changes less than 4% over the temperature range, indicating that the interaction of the benzoate ion with the solvent changes little if at all with increasing temperature. The pK a(m) vs.T data show that ΔH o decreases quadratically while ΔC p o increases linearly withT although, over the 75°C range, ΔC p o increases only about 6 cal-mole?1 deg?1 around an average of ?37 cal-mole?1deg?1. The acid dissociation constants as derived from the conductance-molal concentration analysis show an average uncertainty of about 0.1% and are fitted to within about 0.01% by the equation $$p{\text{K}}_{\text{a}} (m) = - 75.5422 + 3136.34/T + 28.7965 log T - 6.8139 {\text{x}} 10^{ - 3{\text{T}}} $$ whereT is the absolute temperature.  相似文献   

5.
Unlike linear extracting agents, in the extraction of metal salts from aqueous solutions of inorganic acids with crown ethers, the inclusion compounds, whose composition depends on several external and internal factors, go to the organic phase. The study of the molecular structure of the formed complexes by X-ray diffraction analysis showed that adducts of crown ethers with inorganic acids are host–guest complexes in which the hydroxonium ion is in the polyether macrocycle cavity. When the aqueous phase contains metal ions capable of displacing the hydroxonium ions from the macrocycle (K+, Pb2+, Hg2+, Sr2+, NH4 +), complexes containing metal cations as the guest in the macrocycle cavity, according to X-ray diffraction data, go to the organic phase. In addition, metals forming ionic associates (AuCl4 -, FeCl4 -, GaCl4 -) in an aqueous solution are extracted with crown ethers in accordance with the anion-exchange mechanism. A system in which traces of metals in the 2 M HNO3 +5 M HCl mixture serve as the aqueous phase was proposed for estimation of the general extraction ability of crown ethers. Such a system can be used for metal extraction via any possible mechanism. The stereochemical peculiarities of the extraction ability of crown ethers (compared to linear molecules) can be used for selective extraction and separation of metals.  相似文献   

6.
Isopiestic vapor pressure measurements have been used to determine the osmotic coefficients of aqueous solutions of iodic acid at molalities from 0.1 to 17 mole-kg?1 at 25°C. The isopiestic standards were solutions of sodium chloride and solutions of sulfuric acid. Because of the corrosive nature of iodic acid, platinum cups were used. Stoichiometric activity cofficients of iodic acid were derived by a Gibbs-Duhem integration. The activity coefficients for solutions of molality greater than 0.5 mole-kg?1 cannot be accounted for in terms of the two equilibria, namely, the acidic dissociation of iodic acid and formation of the ion H(IO3) 2 ? , shown by Pethybridge and Prue to explain adequately the behavior in dilute solutions. The activity coefficient is unexpectedly small in concentrated solutions, suggesting the formation of neutral aggregates of iodic acid. The presence of dimers and tetramers, or alternatively trimers and tetramers, can explain the observed results up to a molality of 7 mole-kg?1.  相似文献   

7.
A novel electrochemical method based on controlled-potential electrolysis has been developed for the elucidation of the ion transfer at the interface between two immiscible electrolyte solutions (ITIES). A relationship between the applied interfacial potential (Eapp) and the amount of the ion transferred (Atr) was investigated after an electrolytic equilibrium was attained by controlled-potential electrolysis. The Atr was determined chemically or radiometrically instead of by current measurement. It was found that (i) controlled-potential electrolysis was applicable to the study of the transfer of such hydrophilic ions as transition metal ions which gave no appreciable current within the potential window in voltammetry or polarography at ITIES, (ii) controlled-potential electrolysis in combination with a sensitive analytical method enabled a study of the transfer reaction of an ion of very dilute concentration, and (iii) even when the transfer reaction of an ion was irreversible or quasi-reversible, a standard ion transfer potential could be determined by controlled-potential electrolysis without using a kinetic parameter. The controlled-potential electrolysis method developed was applied to the transfer reactions of actinide ions such as UO2 2+ and Am3+ from aqueous solution to nitrobenzene solution in the absence or presence of an ionophore facilitating the transfer. The Gibbs energy for the transfer of actinide ion and a stability constant of the complex between an actinide ion and the ionophore in nitrobenzene solution were determined from log D versus Eapp plots (D the ratio of the concentration of the ion in nitrobenzene solution to that in aqueous solution). The feasibility of controlled-potential electrolysis as a method for electrolytic separation of actinide ions is discussed.  相似文献   

8.
Adsorption of Zn2+ and Cd2+ ions from aqueous waste solutions on iron(III) titanate as inorganic ion exchange material was investigated to determine the effect of contact time, pH of solution and the reaction temperatures. Batch kinetic studies were carried out and showed that the time of equilibrium for both Zn2+ and Cd2+ ions was attained within three hours, and the order of kinetic reaction is the first order reaction. Batch distribution coefficients of Zn2+ and Cd2+ ions on iron(III) titanate as a function of pH have been studied at 25, 40 and 60 ± 1°C. From the obtained results we found that the K d values decreased with increasing reaction temperatures. Enthalpy change (H) values for Zn2+ and Cd2+ ions were found to be –8.19 and –22.49 kJ/mol, respectively. The data of adsorption of Zn2+ and Cd2+ ions at various concentrations were fitted with the Freundlich isotherm. Finally, separation of the above mentioned cations on iron(III) titanate in a column was performed.  相似文献   

9.
Electrical conductivities of dilute aqueous solutions of aluminum sulfate were determined and analyzed in terms of a strongly associated electrolyte of the 3:2 type. The conductivities reported here were determined from 15 to 35 °C. Representation of conductances, in the framework of the ion association model, was performed using the Quint–Viallard conductivity equations for highly charged electrolytes and the Debye–Hückel expression for activity coefficients. Determined apparent association constants K a(T) were considered as adjustable parameters. The determined limiting conductances of the trivalent aluminum ion λ0((1/3)Al3+) are considerably higher than those reported in the literature. Available specific conductivities in concentrated aqueous solutions of aluminum sulfate were fitted by a new empirical equation with only three adjustable parameters.  相似文献   

10.
The limiting conductivity of Dy3+(aq) has been determined for the first time by linear extrapolation of conductivity measured in dilute aqueous solutions of (DyCl3+HCl) at 25°C as o(Dy3+, aq) = (62.9±0.7) S-cm2-eq-1. A second set of conductivity measurements in dilute aqueous solutions of DyCl3 has given evidence of very slight hydrolysis of the cation, with a first hydrolysis constant of 6 x 10–8 mol-dm–3 (pK=7.2±0.5) calculated by applying the Onsager-Kim law of electrolyte mixtures.  相似文献   

11.
Aqueous acidified solutions of the rare-earth-element (REE) triflates (Gd(CF3SO3)3(aq), Dy(CF3SO3)3(aq), Nd(CF3SO3)3(aq), Er(CF3SO3)3(aq), Yb(CF3SO3)3(aq) and Y(CF3SO3)3(aq)) have been prepared by the dissolution of the corresponding REE oxides in dilute aqueous trifluoromethanesulfonic acid (triflic acid, CF3SO3H(aq)). Relative densities and relative massic heat capacities have been measured for these systems over the approximate ionic strength range 0.10≤I/(mol?kg?1)≤1.35 at T=(288.15, 298.15, 313.15 and 328.15) K and p=0.1 MPa. These measurements were completed using a Sodev O2D vibrating tube densimeter and Picker-flow microcalorimeter, respectively. Relative densities and relative massic heat capacities for aqueous solutions of triflic acid and its sodium salt have also been measured over the concentration range 0.018≤m 2/(mol?kg?1)≤0.23 over the same temperature range at p=0.1 MPa. Young’s rule has been used to calculate apparent molar volumes and apparent molar heat capacities of the aqueous solutions of REE triflate salts from the calculated apparent molar properties of the acidified salt solutions. These properties have been modeled using the Pitzer ion-interaction equations. The apparent molar properties of aqueous triflic acid solutions and aqueous solutions of its sodium salt have also been modeled using the same Pitzer ion-interaction equations. The apparent molar properties at infinite dilution obtained from our property modeling have been used to calculate single ion volumes and single ion heat capacities for each of the aqueous ions; Gd (aq) 3+ , Dy (aq) 3+ , Nd (aq) 3+ , Er (aq) 3+ , Yb (aq) 3+ , and Y (aq) 3+ . The reported single ion values have been compared with those previously reported in the literature.  相似文献   

12.
The atomic structures, and thereby the coordination chemistry, of metal ions in aqueous solution represent a cornerstone of chemistry, since they provide first steps in rationalizing generally observed chemical information. However, accurate structural information about metal ion solution species is often surprisingly scarce. Here, the atomic structures of Ga3+ ion complexes were determined directly in aqueous solutions across a wide range of pH, counter anions and concentrations by X-ray pair distribution function analysis and 71Ga NMR. At low pH (<2) octahedrally coordinated gallium dominates as either monomers with a high degree of solvent ordering or as Ga-dimers. At slightly higher pH (pH ≈ 2–3) a polyoxogallate structure is identified as either Ga30 or Ga32 in contradiction with the previously proposed Ga13 Keggin structures. At neutral and slightly higher pH nanosized GaOOH particles form, whereas for pH > 12 tetrahedrally coordinated gallium ions surrounded by ordered solvent are observed. The effects of varying either the concentration or counter anion were minimal. The present study provides the first comprehensive structural exploration of the aqueous chemistry of Ga3+ ions with atomic resolution, which is relevant for both semiconductor fabrication and medical applications.

With changing pH four different structural regions in Ga3+ aqueous solutions are observed. In contrast the effects of different anions and concentrations are minimal.  相似文献   

13.
A solution of cobalt(II) chloride in HCI is commonly used to examine various effects, such as changes in temperature and concentration, on the Co(H2O)62+/CoCl42– equilibrium (Lc Chaâtcliers principle). In aqucous solution the cobalt(II) ion exists as a mixture of two complex ions at equilibrium, blue CoCl42– and pink Co(H2O)62+. As the ions have different colors, it is easy to determine the position of the equilibrium. A series of experiments was designed to allow students to examine the concentration of chloride ion, the dehydration effect, and how dissociation effects the equilibrium. Different compounds were added to the aqueous solution of cobalt(II) ion, and the position of the equilibrium was determined either visually or, more quantitatively, by means of UV–vis spectroscopy. This exercise is suitable for general chemistry students and is designed to introduce them to the complexity of the actual chemical reaction rather than presenting them with a simplistic model.  相似文献   

14.
Isotherms of copper cation sorption by H-ZSM-5 zeolite from aqueous and aqueous ammonia solutions of copper acetate, chloride, nitrate, and sulfate are considered in terms of Langmuir’s monomolecular adsorption model. Using UV-Vis diffuse reflectance spectroscopy, IR spectroscopy, and temperatureprogrammed reduction with hydrogen and carbon monoxide, it has been demonstrated that the electronic state of the copper ions is determined by the ion exchange and heat treatment conditions. The state of the copper ions has an effect on the redox properties and reactivity of the Cu-ZSM-5 catalysts in the selective catalytic reduction (SCR) of NO with propane and in N2O decomposition. The amount of Cu2+ that is sorbed by zeolite H-ZSM-5 from aqueous solution and is stabilized as isolated Cu2+ cations in cationexchange sites of the zeolite depends largely on the copper salt anion. The quantity of Cu(II) cations sorbed from aqueous solutions of copper salts of strong acids is smaller than the quantity of the same cations sorbed from the copper acetate solution. When copper chloride or sulfate is used, the zeolite is modified by the chloride or sulfate anion. Because of the presence of these anions, the redox properties and nitrogen oxides removal (DeNO x ) efficiency of the Cu-ZSM-5 catalysts prepared using the copper salts of strong acids are worse than the same characteristics of the sample prepared using the copper acetate solution. The addition of ammonia to the aqueous solutions of copper salts diminishes the copper salt anion effect on the amount of Cu(II) sorbed from these solutions and hampers the nonspecific sorption of anions on the zeolite surface. As a consequence, the redox and DeNO x properties of Cu-ZSM-5 depend considerably on the NH4OH/Cu2+ ratio in the solution used in ion exchange. The aqueous ammonia solutions of the copper salts with NH4OH/Cu2+ = 6–10 stabilize, in the Cu-ZSM-5 structure, Cu2+ ions bonded with extraframework oxygen, which are more active in DeNO x than isolated Cu2+ ions (which form at NH4OH/Cu2+ = 30) or nanosized CuO particles (which form at NH4OH/Cu2+ = 3).  相似文献   

15.
The adsorption of thorium(IV) ions on molecular sieve (13X type) powder from aqueous solutions has been studied as a function of shaking time pH, thorium ion concentration and temperature. The conditions of maximum adsorption of thorium ions obeys Langmuir and D-R isotherms over the entire concentration range studied. Thermodynamic quantities such as H, G and S have been calculated fromK D values determined at various temperatures. The results show endothermic heat of adsorption, but negative free energy value indicates that the process of thorium adsorption on molecular sieve powder is favored at high temperature. The influence of various cations and anions on thorium(IV) ion adsorption was examined. A wavelength dispersive X-ray fluorescence spectrometer was used for measuring the thorium ion concentration in solutions.  相似文献   

16.
The competitive sorption of Cu(II), Eu(III) and U(VI) ions in aqueous solutions by TiO2, has been investigated by potentiometry at I = 0.1 M NaClO4, 25 °C and under atmospheric conditions. For Cu(II) ions the investigation was performed directly by means of a Cu2+ ion selective electrode, whereas for the Eu(III) and U(VI) ions indirectly based on competition reactions between the Cu(II) ion and the Eu(III) and U(VI) ions. Numerical analysis of the experimental data supports the formation of inner-sphere surface complexes and allows the evaluation of the formation constant of the (TiO)2Cu, which is found to amount to log β* = 4.3 ± 0.4. Addition of competing Eu(III) and U(VI) ions in the aqueous system leads to replacement of the Cu(II) by the competitor metal ion. Evaluation of the potentiometric data obtained from competition experiments indicates on an ion exchange mechanism. The formation constant of the Eu(III) and U(VI) species adsorbed on TiO2 is found to be log β* = 4.4 ± 0.7 and 4.8 ± 0.8, respectively. The relative affinity of the TiO2 surface for the metal ions under investigation is U(VI) > Eu(III) > Cu(II).  相似文献   

17.
18.
The apparent molar volumes, V,2, of glycine, alanine, -amino-n-butyric acid, valine, leucine, and lysine monohydrochloride have been determined in aqueous solutions of 0.05, 0.1, and 0.4 mol-kg–1 Triton X-100 (TX-100), and the partial specific volume, v0, of hen-egg-white lysozyme in 0.4 mol-kg–1 TX-100 by density measurements at 298.15 K. These data have been used to calculate the infinite dilution apparent molar volumes, V2,m0, for the amino acids in aqueous TX-100 solutions and the standard partial molar volumes of transfer, tr V2,m0, of the amino acids from water to the aqueous surfactant solutions. The linear correlation of V2,m0 for a homologous series of amino acids has been utilized to calculate the contribution of the charged end groups (NH3+, COO), CH2 group and other alkyl chains of the amino acids to V2,m0. The results on tr V2,m0, of amino acids from water to aqueous TX-100 solutions have been interpreted in terms of ion–ion, ion–polar, hydrophilic–hydrophilic and hydrophobic–hydrophobic group interactions. For all the six amino acids studied, the values of tr V2,m0 from water to all the studied concentrations of aqueous TX-100 are small in spite of their different hydrophobic content, indicating an overall balance in interactions of zwitterionic/hydrophilic groups of amino acids with the hydrophilic groups of TX-100, and of hydrophobic and ionic/hydrophilic groups of the amino acids with hydrophobic groups of TX-100. Comparison of the interactions of the amino acids with nonionic, anionic and cationic surfactants has also been made and discussed. The partial specific volume of transfer of lysozyme from water to aqueous TX-100 solutions also indicates a balance of the hydrophobic and hydrophilic interactions in the protein–nonionic surfactant system.  相似文献   

19.
The apparent molar volumes, V,2, of glycine, L-alanine, DL--amino-n-butyric acid, L-valine, and L-leucine have been determined in aqueous 0.25, 0.75, 1.0, and 1.5 mol-dm–3 tetraethylammonium bromide (TEAB) solutions by density measurements at 298.15 K. These data have been used to calculate the infinite dilution apparent molar volumes, V2,m, for the amino acids in aqueous tetraethylammonium bromide and the standard partial molar volumes of transfer (tr V2,m) of the amino acids from water to the aqueous salt solutions. The linear correlation of V2,m for a homologous series of amino acids has been utilized to calculate the contribution of the charged end groups (NH3+, COO), CH2 group, and other alkyl chains of the amino acids to V2,m. The results of the standard partial molar volumes of transfer from water to aqueous tetraethylammonium bromide have been interpreted in terms of ion–ion, ion–polar, and hydrophobic–hydrophobic group interactions. The volume of transfer data suggest that ion–ion or ion–hydrophilic interactions are predominant in the case of glycine and alanine, and hydrophobic–hydrophobic group interactions are predominant in the case of DL--amino butyric acid, L-valine, and L-leucine.  相似文献   

20.
Thermodynamics of aqueous solutions of the alkali metal sulfates   总被引:1,自引:0,他引:1  
The available thermodynamic properties for aqueous solutions of each of the alkali metal sulfates have been combined and analyzed within the framework of the ion interaction model at temperatures up to 225°C. It was necessary to set 1 equal to 1.4kg1/2-mol–1/2 in order to obtain a satisfactory fit. The temperature dependence of the ion interaction parameters was given the functional form used by Rogers and Pitzer(1) in their study of Na2SO4(aq). With few exceptions, it was possible to reproduce the available thermodynamic data for aqueous solutions of the alkali metal to within the estimated experimental error. Thermodynamic results for Na2SO4(aq) appear to be adequate in this temperature range, but enthalpy and heat capacity data for the other alkali metal sulfate solutions are conspicuously lacking. Activity coefficients of these electrolytes decreased to less than 0.1 at moderate molalities at the higher temperatures, and their order changed with increasing temperature; two results which could be due to a combination of hydration and association effects.Research sponsored by the Division of Chemical Sciences, Office of Basic Energy Sciences of the U.S. Department of Energy under contract DE-AC05-840R21400 with the Martin Marietta Energy Systems, Inc.  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