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 comprehensive thermodynamic model based on the electrolyte NRTL (eNRTL) activity coefficient equation is developed for the NaCl + H2O binary, the Na2SO4 + H2O binary and the NaCl + Na2SO4 + H2O ternary. The NRTL binary parameters for pairs H2O-(Na+, Cl) and H2O-(Na+, SO42−), and the aqueous phase infinite dilution heat capacity parameters for ions Cl and SO42− are regressed from fitting experimental data on mean ionic activity coefficient, heat capacity, liquid enthalpy and dissolution enthalpy for the NaCl + H2O binary and the Na2SO4 + H2O binary with electrolyte concentrations up to saturation and temperature up to 473.15 K. The Gibbs energy of formation, enthalpy of formation and heat capacity parameters for solids NaCl(s), NaCl·2H2O(s), Na2SO4(s) and Na2SO4·10H2O(s) are obtained by fitting experimental data on solubilities of NaCl and Na2SO4 in water. The NRTL binary parameters for the (Na+, Cl)-(Na+, SO42−) pair are regressed from fitting experimental data on dissolution enthalpies and solubilities for the NaCl + Na2SO4 + H2O ternary.  相似文献   

2.
In the system Na3PO4Na2SO4, the high-temperature, cubic γ form of Na3PO4 forms an extensive range of solid solutions: Na3−x(P1−xSx)O4, 0 < x < (0.57 to 0.70, depending on temperature). For compositions in the range x = ca. 0.33 to 0.57, these γ solid solutions are thermodynamically stable at all temperatures. The conductivity of the γ solid solutions increases with increasing x and reaches a maximum at x = 0.5 to 0.6, with values of 2 × 10−5 ohm−1 cm−1 at 100°C, rising to 1.3 × 10−2 ohm−1 cm−1 by 300°C; this conductivity increase with x is attributed to an increase in the sodium ion vacancy concentration, associated with the solid solution mechanism Na + PS. The phase diagram for the system Na3PO4Na2SO4 is given together with lattice parameters of the γ solid solutions.  相似文献   

3.
The Fourier transform infrared-attenuated total reflectance (FTIR-ATR) difference spectra of aqueous MgSO4, Na2SO4, NaCl and MgCl2 solutions against pure water were obtained at various concentrations. The difference spectra of the solutions showed distinct positive bands and negative bands in the O–H stretching region, indicating the influences of salts on structures of hydrogen-bonds between water molecules. Furthermore the difference spectra of MgCl2 solutions against NaCl solutions and those of MgSO4 solutions against Na2SO4 solutions with the same concentrations of anions (Cl? or SO 4 2? , respectively) allowed extracting the structural difference of the first hydration layer between Mg2+ and Na+. Using SO 4 2? as a reference ion, structural information of the first hydration layer of the Cl? anion was obtained according to the difference spectra of MgCl2 solutions against MgSO4 solutions and those of NaCl solutions against Na2SO4 solutions containing the same concentrations of cations (Mg2+ or Na+, respectively). The positive peak at ~3,407 cm?1 and negative peak at ~3,168 cm?1 in these spectra indicated that adding Cl? decreased the strongest hydrogen-bond component and increased the relatively weaker one.  相似文献   

4.
When the sodium ion (Na+) concentration is increased above 0.5 mol-dm−3 (M), the concentrations of dissolved silica in aqueous sodium chloride (NaCl) and sodium nitrate (NaNO3) solutions decrease because of the salting out effect. On the other hand, the concentration of the dissolved silica in aqueous sodium sulfate (Na2SO4) solutions increases monotonously as the concentration of Na+ is increased above 0.5 M. The purpose of this study is to determine the reasons why the salting-out effect is not observed in Na2SO4 solutions. FAB-MS (Fast Atom Bombardment Mass Spectrometry) was used to sample directly the silica species dissolved in aqueous Na2SO4, NaCl, and NaNO3 solutions. In the FAB-MS spectra of these solutions, the peak intensity ratios of the linear tetramer to the cyclic tetramer largely increased for Na+ concentrations between (0.1 and 1) M. This shows that some characteristics of the Na2SO4 solutions are similar to those of the NaCl and NaNO3 solutions. In Na2SO4 solutions, however, when the concentration of Na+ is higher than 1 M, the peak intensity of the dimer is much higher than those of the other silicate complexes. In Na2SO4 solutions, the SO42− ion undergoes partial hydrolysis to form HSO4 and OH is produced. In particular, in the range where the concentration of SO42− is high, the pH of the solution increases slightly. This higher pH yields more dimers from the hydrolysis of silicate complexes. This increase in dimer production agrees with the observation that silica dissolves in sodium hydroxide (NaOH) solutions mainly as a dimer when the concentration of NaOH is less than 0.1 M. In Na2SO4 solutions at high concentrations, a salting-out effect is not observed for silica. This is due to the increase in the concentration of OH, which accelerates the hydrolysis of silica and results in dimer formation.  相似文献   

5.
研究了LiZr2(PO4)3在水溶液中的Na/Li和Ag/Li离子交换行为.结果表明,LiZr2(PO4)3对Na+和Ag+离子均具有很高的选择性,且对Ag+的选择性高于Na+.LiZr2(PO4)3与Ag+的离子交换反应是通过形成固溶体的形式进行的,而与Na+的离子交换反应则是通过置换进行的.温度升高有利于提高LiZr2(PO4)3上Na/Li和Ag/Li的离子交换反应速度.  相似文献   

6.
A complete, critical evaluation of all phase diagram and thermodynamic data was performed for all phases of the (Na2SO4 + K2SO4 + Na2S2O7 + K2S2O7) system and optimized model parameters were obtained. The Modified Quasichemical Model in the Quadruplet Approximation was used for modelling the liquid phase. The model evaluates first- and second-nearest-neighbour short-range ordering, where the cations (Na+ and K+) are assumed to mix on a cationic sublattice, while anions were assumed to mix on an anionic sublattice. The Compound Energy Formalism was used for modelling the solid solutions of (Na,K)2SO4 and (Na,K)2S2O7. The models can be used to predict the thermodynamic properties and phase equilibria in multicomponent heterogeneous systems. The experimental data from the literature were reproduced within experimental error limits.  相似文献   

7.
The kinetics of the Fe3+/Fe2+ reaction on a Pt rotating disc electrode was studied in solutions of 0.5 M H2SO4 and 0.5 M Na2SO4 (pH 2.2). Taking into account formation of sulphate complexes the conclusion was made that the main contribution to the reaction rate is due to FeSO4+ and FeSO4 complexes. Extended Tafel plots obtained by Randles analysis from experimental current-voltage curves were corrected for the 2 potential. The latter was evaluated according to the Gouy-Chapman theory by using the surface charge density values deduced from thermodynamic theory and measurements of other authors. Tafel plots were approximated by parabolas and the reorganization energy was calculated as 33 kJ mol?1 and 51 kJ mol?1 for Fe3+/Fe2+ in H2SO4 and Na2SO4, respectively. The comparison of these values with theoretically predicted ones was made. From the magnitude of the pre-exponential factor of the true rate constant it was concluded that the Fe3+/Fe2+ electron transfer reaction is non-adiabatic in nature.  相似文献   

8.
We have measured the ionic conductivities of pressed pellets of the layered compounds MUO2PO4 · nH2O, and correlated the results with TGA data. The conductivities (in ohm?1 m?1), at temperatures increasing with decreasing water content over the range 20 to 200°C, were approximately as follows: Li+4H2O, 10?4; Li+, Na+, K+, and NH4+3H2O, 10?4, 10?2, 10?4, and 10?4; H+, Li+, and Na+1.5H2O, 10?2, 10?4, and 10?4; Na+1H2O, 10?5; H+, K+, and NH4+0.5H2O, all 10?5; and H+, Li+, Na+, K+, NH+4, and 12Ca2+OH2O, 10?5, 10?5, 10?4, 10?5, 10?5, and 10?6. A ring mechanism is proposed to account for the high conductivity found in NaUO2PO4 · 3.1H2O. The accurate TGA data showed that most of the hydrates had water vacancies of the Schottky type, and should be represented as MUO2PO4(A ? x)H2O, where x can be between 0 and 0.3.  相似文献   

9.
The kinetics of the redox reaction between mandelic acid (MA) and ceric sulfate have been studied in aqueous sulfuric acid solutions and in H2SO4? MClO4 (M+ = H+, Li+, Na+) and H2SO4? MHSO4 (M+ = Li+, Na+, K+) mixtures under various experimental conditions of total electrolyte concentration (that is, ionic strength) and temperature. The oxidation reaction has been found to occur via two paths according to the following rate law: rate = k[MA] [Ce(IV)], where k = k1 + k2/(1 + a)2[HSO4?]2 = k1 + k2/(1 + 1/a)2[SO42?]2, a being a constant. The cations considered exhibit negative specific effects upon the overall oxidation rate following the order H+ ? Li+ < Na+ < K+. The observed negative cation effects on the rate constant k1 are in the order Na+ < Li+ < H+, whereas the order is in reverse for k2, namely, H+ ? Li+ < Na+. Lithium and hydrogen ions exhibit similar medium effects only when relatively small amounts of electrolytes are replaced. The type of the cation used does not affect significantly the activation parameters.  相似文献   

10.
11.
From extraction experiments and γ-activity measurements, the exchange extraction constants corresponding to the general equilibrium M+(aq)+NaL+(nb)⇔ML+(nb)+Na+(aq) taking place in the two-phase water-nitrobenzene system (M+ = Li+, H3O+, NH4+, Ag+; L = hexaethyl calix[6]arene hexaacetate; aq = aqueous phase, nb = nitrobenzene phase) were determined. Furthermore, the stability constants of the ML+ complexes in water saturated nitrobenzene were calculated; they were found to increase in the cation order H3O+<NH4+<Li+<Ag+.  相似文献   

12.
(1 ? x)AgPO3xAg2SO4 homogeneous glasses obtained by quenching of a melt of the two salts are pure ionic Ag+ conductors. The RT conductivity is increased from 2.5 × 10?7 to 4 × 10?6 (Ω cm)?1 when the ratio of Ag2SO4 is increased from 0 to 0.3. Raman spectroscopy shows that no modifications of the (PO3) chain skeleton occur by adding Ag2SO4. The low-frequency Raman band lying at about 55 cm?1 is quantitatively correlated to Ag+ oscillations, the hopping distance decreasing from 3.0 to 2.7 Å if a jump process between regular Ag+ sites is considered.  相似文献   

13.
H.F. Rexroat  N.S. Rowan 《Polyhedron》1985,4(8):1357-1363
trans-[Co(en)2(SO3)(H2O)]+ reacts with imidazole (ImH) and imidazole containing ligands (L) to form trans-[Co(en)2(SO3)L]+ in the pH range 6.0–9.0. The complex seems to react both in the hydroxy and in the aquo form. The rate constant for the reaction of imidazole with the aquo form is 6.0±0.7 and 4±1M?1s?1 for the reaction with the hydroxy form at 25°C. The apparent equilibrium constant for formation of the imidazole complex at pH 7 is consistent with the value of 3 x 102 measured previously. Appreciable amounts of complex form only in the pH 6–9 range. Above pH 9 NMR spectra show that even the immediate products are different. In aged solutions at all pHs other products form.  相似文献   

14.
The system POCl3–NaAlCl4 was investigated by measuring the conductivity and the Raman and NMR spectra (27Al, 23Na and 31P) as a function of the mol fraction x of NaAlCl4 in POCl3. Additionally, Raman spectra of POCl3 solutions of NaFeCl4, LiAlCl4, LiFeCl4, and KAlCl4 were recorded. In solutions containing Li+ or Na+ ions a liquid to solid (or jelly) phase transition was observed under certain conditions, dependent on salt concentration and temperature. Observed changes in the Raman spectra of the electrolyte solutions in comparison to the pure solvent POCl3 demonstrate the existence of interactions. Clearly, the POCl3 eigenfrequencies and hence the molecules are pertubed. The formation of [M(POCl3)4]+ complexes (M = Li, Na) can be deduced from the Raman measurements. NMR investigations support this conclusion. For assigning of Raman spectra, (Li+, K+) cation and ([FeCl4]?, [SbCl6]?) anion substitutions were employed.  相似文献   

15.
Two ranges of solid solutions were prepared in the system Li4SiO4Li3VO4: Li4?xSi1?xVxO4, 0 < x ? 0.37 with the Li4SiO4 structure and Li3+yV1?ySiyO4, 0.18 ? y ? 0.53 with a γ structure. The conductivity of both solid solutions is much higher than that of the end members and passes through a maximum at ~40Li4SiO4 · 60Li3VO4 with values of ~1 × 10?5 ohm?1 cm?1 at 20°C, rising to ~4 × 10?2 ohm?1 cm?1 at 300°C. These conductivities are several times higher than in the corresponding Li4SiO4Li3(P,As)O4 systems, especially at room temperature. The solid solutions are easy to prepare, are stable in air, and maintain their conductivity with time. The mechanism of conduction is discussed in terms of the random-walk equation for conductivity and the significance of the term c(1 ? c) in the preexponential factor is assessed. Data for the three systems Li4SiO4Li3YO4 (Y = P, As. V) are compared.  相似文献   

16.
Desorption/ionization on silicon (DIOS) mass spectra of model ionic dyes methylene blue (MB+Cl?) and methyl orange (Na+MO?) were studied using p+ type‐derived porous silicon (PS) free layers. As‐prepared PS (PS‐H), the PS thermally oxidized at 300 °C (PS‐OX), PS with chemically grafted cation‐exchanging alkylsulfonic acid (PS‐SO3H) and anion‐exchanging propyl‐octadecyldimethylammonium chloride (PS‐ODMA+Cl?) groups was tested as ionization platforms. Two mechanisms of the methylene blue desorption/ionization were found: (1) the formation of [MB + H]+? ion due to the reduction/protonation of MB+, which is predominant for PS‐H and PS‐OX platforms and (2) direct thermal desorption of the MB+ cation, prevailing for PS‐SO3H. The fragmentation of the cation is significantly suppressed in the latter case. The samples of PS‐SO3H and PS‐ODMA+ Cl? efficiently adsorb the dyes of the opposite charge from their solutions via the ion‐exchange. Consequent DIOS MS studies allow to detect only low fragmented ions (MB+ and MO?, respectively), demonstrating the potential of the ion‐exchange adsorption combined with DIOS MS for the analysis of ionic organic compounds in solutions. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

17.
The transference of water that results from ion migration through the nickel hydroxide precipitate membrane was studied in chloride, perchlorate, nitrate, and sulphate solutions to estimate the transference number of water and the co-ion transport. In the systems of univalent anions, the moles of water transported per mole of electrons in 0.1 N solutions is almost identical to the hydration number of each anion. This water flow decreases gradually as the concentration of external solution increases, because of increase in the co-ion (cation) transport with increasing concentration of the solution. In the system of sulphate solutions the co-ion transport is remarkable, the transport number of Na+ ions being 0.03 in 0.01 N, 0.27 in 0.10 N, and 0.50 in 0.5 N Na2SO4 solution. This large co-ion transport in Na2SO4 solution is attributed to the partical replacement of hydroxyl groups on the membrane by SO2?4 ions, which then acts as a negative fixed charge. The order of the selectivity for co-ion transport is K+ > Na+ > Li+ > Ni2+ ? Mg2+ in sulphate solutions and also in chloride solutions, although the transport number of the cations is much smaller in chloride solution than in sulphate solution.  相似文献   

18.
Phase relations at 1050°C have been determined for M-phase solid solutions in the LiO0.5-NbO2.5-TiO2 ternary phase system by the quench method. Rietveld analysis has been used to help determine phase boundaries and to study structure composition relations. The M-phases have trigonal structures based on intergrowth of corundum-like layers, [Ti2O3]2+, with slabs of (N−1) layers of LiNbO3-type parallel to (0001). Ideal compositions are defined along the pseudobinary join LiNbO3-Li4Ti5O12 by the homologous series formula LiNNbN−4Ti5O3N, N?4. Homologues with N?10 lie to the low-lithia side of the LiNbO3-Li4Ti5O12 join and show extended single-phase solid solution ranges separated by two-phase regions. The composition variations along the solid solutions are controlled by a major substitution mechanism, Li++3Nb5+↔4Ti4+, coupled with a minor substitution 4Li+↔Ti4++3□, where □=vacancy. The latter substitution results in increasing deviations from the stoichiometric compositions A2N+1O3N with increasing Ti substitution. The non-stoichiometry can be reduced by re-equilibration at lower temperatures. Expressions have been developed to describe the compositional changes along the solid solutions.  相似文献   

19.
The rate of oxidation of Ge(II) chloride by large excess of ClO2? ions in HCl, NaCl and Na2SO4 mixed solutions was polarographically observed at various H2O+ and Cl? ion concentrations. The observed rate constant, kobs, is expressed by ko=Kobs/(ClO3?)={k1,(H+)+k2K1(Cl?)2+ K3K2(SO42?)} (H+)/{(H+)1+K1(Cl-)2 +K2(SO42?)} for the following reaction processes, The values were obtained aa k1=1.5410-3liter2 mole2? sec-1, k2=5.00×10-2liter2 mole2? sec-2 and k2=4.30×10-3liter2 mole2? sec-2, K1=1.80× 10-2, K2= 2.43×10-2 mole liter-1 at constant ionic strength I=0.50 M at 30°C.  相似文献   

20.
A Na3V2(PO4)3 sample coated uniformly with a layer of 6 nm carbon has been successfully synthesized by a one-step solid state reaction. This material shows two flat voltage plateaus at 3.4 V vs. Na+/Na and 1.63 V vs. Na+/Na in a nonaqueous sodium cell. When the Na3V2(PO4)3/C sample is tested as a cathode in a voltage range of 2.7-3.8 V vs. Na+/Na, its initial charge and discharge capacities are 98.6 and 93 mAh/g. The capacity retention of 99% can be achieved after 10 cycles. The electrode shows good cycle performance and moderate rate performance. When it is tested as an anode in a voltage range of 1.0-3.0 V vs. Na+/Na, the initial reversible capacity is 66.3 mAh/g and the capacity of 59 mAh/g can be maintained after 50 cycles. These preliminary results indicate that Na3V2(PO4)3/C is a new promising material for sodium ion batteries.  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