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 203 毫秒
1.
Raman OD stretching spectra of alcoholic LiX (X = Cl, Br, I, ClO4 , NO3 , and CH3COO) solutions (alcohols = methanol and ethanol) were measured in the liquid state at room temperature and in the glassy state at liquid nitrogen temperature. The effects of the anions on the Raman OD stretching spectra in these alcoholic solutions are investigated and the structural changes of the solutions are discussed. It is shown that the structure-breaking effects of anions on the intrinsic alcoholic structure increase in the order: Cl < Br < I < ClO4 . From spectral changes, it seems that CH3COOLi exerts little effect on the liquid structure of the alcohol.  相似文献   

2.
Vapor pressure data for solutions of various alkali metal and tetraalkylammonium salts in methanol at 25° and NaI solutions in ethanol, 2-propanol, and acetonitrile at 25° in the concentration range 0.04–1]<0.75 are used to show the applicability of the HNC integral equation method to nonaqueous electrolyte solutions. Friedman's model of overlapping cospheres is used to express the non-electrostatic part of the ion-interaction potential. Data analysis is based on the Rasaiah-Friedman algorithm for the calculation of g++ and g+– functions. After conversion from Lewis-Randall to the McMillan-Mayer system the measured osmotic coefficients of all electrolyte solutions can be reproduced with the help of the calculated correlation functions.  相似文献   

3.
Excess molar volumes VmE of the binary mixtures of (trifluoroethanol + 1-propanol), (trifluoroethanol + 2-propanol), (acetone + water), (methanol + water), (ethanol + water), (1-propanol + water), (2-propanol + water), and the ternary mixtures of (trifluoroethanol + methanol + water), (trifluoroethanol + ethanol + water), (trifluoroethanol~+ 1-propanol + water), (trifluoroethanol + 2-propanol + water) and (trifluoroethanol + acetone + water) were measured with a vibrating tube densimeter at the temperature of 298.15 K and the pressure 101 kPa. The extrema in VmE of trifluoroethanol mixtures occur at –0.690 cm3-mol–1 for (trifluoroethanol + 1-propanol), at –0.990~cm3-mol–1 for (trifluoroethanol + 2-propanol); at 0.562 and –0.973 cm3-mol–1 for (trifluoroethanol + methanol + water), at 0.629 and –0.973 cm3-mol–1 for (trifluoroethanol + ethanol + water), at 1.082 and –0.659 cm3-mol–1 for (trifluoroethanol~+ 1-propanol + water), at 0.998 and –0.991 cm3-mol–1 for (trifluoroethanol~+ 2-propanol + water), and at 0.515 and –1.472 cm3-mol–1 for (trifluoroethanol + acetone + water). The experimental ternary VmE values were predicted by empirical expressions using binary solution data.  相似文献   

4.
Raman spectroscopic measurements were carried out for anhydrous alcoholic rare earth chloride solutions (LnCl3 · 20ROH; ROH = MeOH, EtOH, and n-PrOH) in the glassy state at liquid nitrogen temperature. Series behavior of the Ln–Cl stretching Raman band is examined in conjunction with the formation of chloro complexes. The results are summarized as follows: (1) Comparing the results with those in the liquid state at room temperature, more Cl- ions tend to coordinate to heavy rare earth ions in the three alcohol solutions upon going from a liquid to a glassy state, while the coordination number of the Cl- ions of the light rare earth ions remains almost unchanged on vitrification. (2) In the former half region of the rare earth series, the coordination number of the Cl- ions of the rare earth ions in the methanol LnCl3 solution is apparently the same as that in the ethanol LnCl3 solution. (3) In the latter half region, the higher chloro complexes, such as [LnClx+1(ROH)y-2](z-1) (x + y = 8; x = 1, z = 2 or x = 2, z = 1) are more abundant in the methanol LnCl3 solution than in the ethanol and n-propanol LnCl3 solutions.  相似文献   

5.
A Raman spectral study of 14 solutions of varying bromide to zinc ratios was conducted up to 300°C and 9 MPa. The tetra-, tri-, di- as well as the mono-bromozinc complexes were identified. The signal from the ZnBr+ complex increased in intensity as temperature increased, for solutions of low bromide- to-zinc ratios. The ZnBr 4 2– species was favored at higher Br/Zn ratios, and higher temperatures favored the formation of the species ZnBr2 and ZnBr+ at the expense of ZnBr 4 2– and ZnBr 3 . Although solvated water is probably present in these zinc-bromo complexes, we found no evidence of O–Zn vibrations other than for Zn(H2O) 6 2+ . However, spectra of successive dilutions of solutions with high bromide to zinc ratios show a relative change in species populations thereby suggesting that water activity plays a decisive role in complex formation. For the first time trifluoromethanesulfonic acid (HTFMS) has been used as an internal standard in Raman spectroscopy. This permitted quantitative measurement of stepwise stability constants.  相似文献   

6.
Electromotive force measurements were carried out on the HCl–ZnCl2–H2O system at constant total ionic strengths of 0.1, 0.2, 0.5, 1.0 and 2.0 mol-kg–1 at 25 and 35°C using a cell consisting of Pt, H2(g, 1 atm)|HCl(mA), ZnCl2(mB)|AgCl/Ag. The data were interpreted by the mixed electrolyte equations of Pitzer and Kim in order to evaluate mixing ion-interaction parameters. The activity coefficients of ZnCl2 and the Gibbs excess free energies of mixing are calculated and presented at I=2.0 mol-kg–1 and compared with similar systems containing transition metal chlorides.  相似文献   

7.
The MFTIR IR spectra of solutions of HCl in methanol were obtained in the 900–4000 cm–1 frequency range. It was found that each proton binds two molecules of methanol. The spectra exhibit intense, continuous absorption (CA) with an intensity coefficient at 2000 cm–1 of 174±10 liter/(mole·cm), which is in agreement with the corresponding coefficient for H5O 2 + . The optical densities of CA are linear functions of the concentration of HCl at 900–1600 cm–1; there is no linearity at higher frequencies for CHCl>4 M, and there are less than two molecules of MeOH for each (MeOH)2H+ ion. The results obtained are in agreement with the model in which CA arises in solutions of strong acids because of the interaction of proton vibrations in a strong symmetric H bond with the vibrations of other groups of the proton disolvate.N. N. Semenov Institute of Chemical Physics, Russian Academy of Sciences, 117977 Moscow. Translated from Izvestiya Akademii Nauk, Seriya Khimicheskaya, No. 10, pp. 2261–2268, October, 1992.  相似文献   

8.
Mutual diffusion coefficients and densities were measured for aqueous ZnCl2–KCl mixtures at 25° by using free-diffusion Rayleigh interferometry and pycnometry, respectively. The ZnCl2 concentrations were fixed at 1.5 mol-dm–3, whereas those of KCl were 0.5, 1.25, 2.0, or 4.0 mol-dm–3. This corresponds to a half charged zinc-chlorine storage battery at various suporting electrolyte concentrations. The main-term coefficient of ZnCl2 only varies by 10% with KCl concentration, whereas that of KCl varies by about 22%. The ZnCl2 cross-term coefficient remains small and positive; in contrast the KCl cross-term coefficient goes through a maximum and is negative at high and low KCl concentrations. At KCl concentrations of 0.5 and 4.0 mol-dm–3, solutions with the KCl c0 are statically and dynamically (diffusively) unstable at the top and bottom of the boundary. Evaluation of the parameters of the non-linear least-squares solution to the diffusion equation is difficult for the 1.25 mol-dm–3 KCl case, since this system has nearly equal eigenvalues in its diffusion coefficient matrix.  相似文献   

9.
Isopiestic vapor pressure measurements were made for {xZnCl2+(1−x)ZnSO4}(aq) solutions with ZnCl2 molality fractions of x=(0,0.3062,0.5730,0.7969, and 1) at the temperature 298.15 K, using KCl(aq) as the reference standard. These measurements cover the water activity range 0.901–0.919≤a w≤0.978. The experimental osmotic coefficients were used to evaluate the parameters of an extended ion-interaction (Pitzer) model for these mixed electrolyte solutions. A similar analysis was made of the available activity data for ZnCl2(aq) at 298.15 K, while assuming the presence of equilibrium amounts of ZnCl+(aq) ion-pairs, to derive the ion-interaction parameters for the hypothetical pure binary electrolytes (Zn2+,2Cl) and (ZnCl+,Cl). These parameters are required for the analysis of the mixture results. Although significant concentrations of higher-order zinc chloride complexes may also be present in these solutions, it was possible to represent the osmotic coefficients accurately by explicitly including only the predominant complex ZnCl+(aq) and the completely dissociated ions. The ionic activity coefficients and osmotic coefficients were calculated over the investigated molality range using the evaluated extended Pitzer model parameters.  相似文献   

10.
Several physicochemical techniques have been utilized to study mixed aqueous solutions of ZnCl2 and LiCl along selected pseudobinary composition lines in the concentration range from 2.8 to 22.2m. Measurements of glasstransition temperature, electrical conductance, viscosity, molar volume, refractive index, and proton chemical shift indicate with varying sensitivity that the behavior of these electrolytes is dominated by chlorozinc complexation. The effects of complexation are most sensitively shown by the PMR spectra. In solutions containing more than 6 moles of water per mole of salt, it is probable that several complexation equilibria occur. When the water content is reduced to less than 6 moles per mole of salt, formation of tetrachlorozincate anions appears almost exclusively to be responsible for the composition variation of the physical properties of the solutions. ZnCl 4 = appears to be an even weaker base than ClO 4 .  相似文献   

11.
Photographic Raman spectra were obtained at shifts to ca. 7000 cm–1 for pure water and for a saturated aqueous solution of NaCl using argon ion laser excitation. Raman spectra were also obtained photoelectrically for H2O and D2O between ca. 2500 and ca. 7000 cm–1 using 248-nm excimer laser excitation and boxcar detection. Overtone and combination assignments are presented for H2O and D2O. The first IR OH-stretching overtone from water occurs 215 cm–1 above the first Raman OH-stretching overtone because the IR overtones are dominated by asymmetric stretching. The second OH-stretching Raman overtone from water is estimated to occur near 10,020 ± 20 cm–1, with 9950 cm–1 as a lower limit.  相似文献   

12.
The Raman spectra of the v1-SO 4 2– band in 0.5–2.5 molar aqueous (NH4)2SO4 and ZnSO4 solutions in the temperature range 25–85°C were studied. The molar scattering coefficient of the v1 band is the same for all forms of sulfate in (NH4)2SO4 and ZnSO4 solutions and is independent of temperature up to 85°C. The v1 band profile is symmetrical in (NH4)2SO4 solutions. In ZnSO4 solutions, a shoulder appears on the high frequency side which increases slightly in intensity with increasing concentration and temperature. This high frequency component is attributed to the formation of the contact ion pair (Zn2+·SO 4 2– ). The enthalpy of formation for the contact ion pair is estimated from the Raman data to be approximately 3 kJ-mol–1 which is in reasonable agreement with measurements by other methods.  相似文献   

13.
Raman spectra have been measured for aqueous ZnSO4 solutions under hydrothermal conditions at steam saturation to 244°C; solubility has been recorded as a function of temperature from 25 to 256°C. The high-temperature Raman spectra contained two polarized bands, which suggest that a second sulfato complex, possibly bidentate, is formed in solution, in addition to the 1:1 zinc(II) sulfato complex, which is the only ion pair identified at lower temperatures. Under hydrothermal conditions, it was possible to observe the hydrolysis of the zinc(II) aquo ion by measuring the relative intensity of bands due to SO 4 2– and HSO 4 according to the equilibrium reaction Zn(OH2)6]2+ + SO 4 2– [Zn(OH2)5OH]+ + HSO 4 The precipitate in equilibrium with the solution at 210°C could be characterized as ZnSO4 · H2O (gunningite) by x-ray diffraction (XRD) and Raman and infrared spectroscopy. At 244°C the equilibrium precipitate could be identified as ZnSO4 (zincosite).  相似文献   

14.
Solutions of zinc salts in water, methanol (MeOH), dimethylformamide (DMF) and binary mixtures of water with the two nonaqueous solvents were studied by zinc-67 NMR measurements. Anhydrous zinc nitrate solutions in DMF and MeOH show upfield, concentration independent, chemical shifts at –27 and –19 ppm, respectively, vs. the aqueous solution standard. Addition of DMF or MeOH to an aqueous solution of a zinc salt results in a diamagnetic shift but for the addition of acetonitrile a paramagnetic shift results. In all cases the signal was broadened very considerably, e.g., in ZnCl2 solution the linewidth increased from 40 to 600 Hz in going from water to 35% aqueous MeOH. Both67Zn and13C NMR failed to show any complexation of Zn2+ ion by crown ethers in aqueous solution. A gradual addition of EDTA, of diaza-18-crown-6 or of tetraazacyclotetradecane resulted in an immediate broadening of the67Zn signal which became undetectable when one equivalent of a ligand was added.  相似文献   

15.
The results of an EXAFS investigation of methanol and ethanol solutions of CoBr2 are reported. The curve fitting analysis gives the coordination numbers and the Co–Br and Co-solvent distances. The results for both 0.2M and 3.8M CoBr2 in ethanol show the existence of [CoBr2(C2H5OH)2] as the dominant species, although a considerable amount of octahedral species also exists in the solutions. In 0.2 and 3.3M CoBr2 methanol solutions, the dominant species are [Co(CH3OH)6]2+ and [CoBr(CH3OH)5]+, respectively.  相似文献   

16.
The liquidus temperature and induction periods were measured for crystallization in a system of calcium nitrate, calcium chloride, and water over a concentration range of 5–20 mole% Ca(II), i.e., R=4–18 [R=moles H2O/moles Ca(II)] and ycl=0–1 [ycl=moles Cl/moles (NO 3 +Cl)]. A ternary phase diagram was constructed, and qualitative dependences of the supercooling at which the solution began to crystallize on the system composition were found. A wide range of stability toward crystallization was found for solutions withR=4–10 and ycl=0–0.7 The relationships between the system stability toward crystallization and the viscosity, glass-transition temperature, and the liquidus temperature are discussed.  相似文献   

17.
Heats of dilution of concentrated aqueous solutions (4.43 moles-kg–1) of FeCl2 were measured at 15, 25, and 35°C. The heat capacities of these concentrated solutions were also measured at the same temperatures. From these data the partial molal heat capacity, C p2 0 (FeCl2, aq, 298.15°K)=–2.56±30 J–°K–1–mole–1, was calculated. The partial molal heat capacity of Fe2+(aq), –2±30 J-°K–1-mole–1, was correlated with the correspondence principle equations of Criss and Cobble.  相似文献   

18.
It was established that in UV irradiation of solutions of TiCl4 in methanol, ethanol, isopropanol, and butanol, alcohol-chloride complexes of titanium(III) are formed. The quantum yields of the formation of coordination compounds of titanium(III) depend on the nature of the alcohol: 0.08 (methanol); 0.13 (ethanol); 0.20 (butanol); 0.22 (isopropanol). As complexes of titanium(III) accumulate in solution, there is a liberation of molecular hydrogen. The quantum yields of the formation of hydrogen, determined in a steady-state process, are correlated with the values of the C-H bond energy at the -carbon atom of the alcohol and are equal to 2·10–3, 3.4·10–3, 4.3·10–3 and 1·10–2 for solutions in methanol, butanol, ethanol, and isopropanol, respectively. A substantial increase in the quantum yield of the formation of molecular hydrogen was detected when a heterogeneous catalyst (palladium on silica gel) was used, and the possible mechanism of the process of photocatalytic liberation of hydrogen from alcohols with the participation of titanium complexes is discussed.Translated from Teoreticheskaya i Éksperimental'naya Khimiya, Vol. 3, No. 2, pp. 181–186, March–April, 1987.The authors would like to thank V. M. Granchak for his participation in the discussion of the results.  相似文献   

19.
The standard potential of the Zn–Hg (sat)/ZnCl2(M)/AgCl/Ag cell was determined at 25°C using several extrapolation procedures and the value 0.9843 V is proposed for E°. The emf of the Zn–Hg (sat)/ZnCl2(M), salt (M)/AgCl/Ag cell [salt=NaClO4, LiClO4, Mg(ClO4)2, Mg(NO3)2] have been measured at different concentrations of salt. From these data, the mean ionic activity coefficients of ZnCl2 are determined and their variations explained with the aid of Pitzer's treatment. It seems necessary to take into account structure in the ionic cospheres in order to explain the observed variations.  相似文献   

20.
Raman and IR data for aqueous CdSO4 and (NH4)2SO4 solutions have been recorded over broad concentration and temperature ranges. Whereas the v1-SO 4 2– band profile is symmetrical in (NH4)2SO4 solutions, in CdSO4 solutions a shoulder appears on the high frequency side which increases in intensity with increasing concentration and temperature. The molar scattering coefficient of the v1-SO 4 2– band is the same for all forms of sulfate in (NH4)2SO4 and CdSO4 solutions and is independent of temperature up to 99°C. The high frequency shoulder is attributed to the formation of a contact ion pair [Cd2+OSO 3 2– ] (11 associate). Also the v3-SO 4 2– antisymmetric stretching mode shows a splitting in the CdSO4 solution. Further spectroscopic evidence for contact ion pair formation is provided by IR spectroscopy. No higher associates or anionic complexes are required to interpret the spectroscopic data. The degree of association has been measured as a function of concentration and temperature. The thermodynamic association constant, KA=0.15±0.05 kg-mol–1 at 25°C is estimated from the Raman data by an extrapolation procedure by taking account of the activity coefficients. Values are reported for the activity coefficient of the ion pair. From the Raman temperature dependence studies, the enthalpy of formation for the contact ion pair is estimated to be 10±1 kJ-mol–1.  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