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 28 毫秒
1.
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   

2.
Raman spectra of aqueous FeSO4 and (NH4)2SO4 solutions have been recorded over broad concentration and temperature ranges. Whereas the v1-SO 4 2- band profile is symmetrical in noncomplexing (NH4)2SO4 solutions, in FeSO4 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 FeSO4 solutions and is independent of temperature up to 150‡C, the highest temperature studied. The high-frequency shoulder is attributed to the formation of a contact ion pair, Fe2+OSO3/2-, as is the splitting of the v3-SO 4 2- antisymmetric stretching mode which is observed in the FeSO4 solution. The bending modes v2-SO 4 2- and v4-SO 4 2- , normally forbidden in the isotropic spectrum, show a gain in intensity with increasing ion-pair formation. A polarized band has been assigned to the Fe2+-O ligand vibration. No higher associates or anionic complexes are required to interpret the spectroscopic data. No evidence of contact ion pairing between Fe2+ and HSO4 4 - could be detected at temperatures up to 303‡C in 1 molal solutions of FeSO4 with an excess of 2 molal H2SO4.  相似文献   

3.
The Raman spectra of N2O4 solutions in organic solvents have been recorded. The frequencies ofv 1,v 2, andv 3 bands of N2O4 increase with increasing solvent electron-donor properties. Especially large changes ofv 3 N-N stretching band have been observed (254.5 cm–1 in n-hexane, 276.5 cm–1 in 1,4-dioxane). The ab initio calculations have shown that the interaction between N2O4 and electron-donor molecules causes an increase of N-N and N-O stretching and O-N-O bending force constants of N2O4 in agreement with the results of Raman study.  相似文献   

4.
Raman and FTIR spectra of guanidinium zinc sulphate [C(NH2)3]2Zn(SO4)2 are recorded and the spectral bands assignment is carried out in terms of the fundamental modes of vibration of the guanidinium cations and sulphate anions. The analysis of the spectrum reveals distorted SO42− tetrahedra with distinct S–O bonds. The distortion of the sulphate tetrahedra is attributed to Zn–O–S–O–Zn bridging in the structure as well as hydrogen bonding. The CN3 group is planar which is expressed in the twofold symmetry along the C–N (1) vector. Spectral studies also reveal the presence of hydrogen bonds in the sample. The vibrational frequencies of [C(NH2)3]2 and HC(NH2)3 are computed using Gaussian 03 with HF/6-31G* as basis set.  相似文献   

5.
14N chemical shifts and linewidths were determined for NO 3 and NH3 in liquid ammonia solutions of thallium nitrate at concentrations between 0.07 and 10 M. The concentration dependences of the14NO 3 shift and linewidth are consistent with the presence of C2v ion pairs at a 2:1 mole ratio of NH3 to TINO3 and C3v ion pairs at mole ratios of 3:1 or higher. Previous studies had indicated the formation of ion pairs at low concentrations. The small value of the14NO 3 linewidth below 1 M suggests that these are contact ion pairs. Studies of the14NH3 linewidth as a function of thallium salt concentration indicate slow solvent exchange at very high concentrations.14NH3 exhibits a downfield shift upon incorporation into the solvation sphere of the Tl+NO 3 ion pair, in constrast to upfield shifts reported for complexation with transition metal cations.  相似文献   

6.
Osmotic coefficients of water have been measured isopiestically for the entire region of homogeneous ternary solutions for the Rb2SO4- (NH4)2SO4-H2O system at 25°C. One might expect that water isoactivity lines should be straight since this system involves a continuous series of solid solutions. The related systems (K2SO4-Rb2SO4-H2O and (K2SO4- (NH4)2 SO4-H2O) obey the linearity of water isoactivity lines rule. Contrary to expectations, the (Rb2SO4-(NH4)2-SO4-H2O appears to be the first water–salt system containing continuous solid solutions in which the mentioned rule is not obeyed.  相似文献   

7.
The relative sound speed of mixtures of aqueous solutions of NaCl–MgSO4 and MgCl2–Na2SO4 at I=0.1 and 0.5m have been determined at 5, 15, and 25°C and pressures to 1000 bars. The resulting sound speeds, adiabatic and apparent molal compressibilities have been compared to results estimated from binary solutions using an additivity principle — Young's rule. The estimated sound speeds agree with the measured values for the NaCl–MgSO4 system to ±0.15 m-sec–1 and for the Na2SO4–MgCL2 system to ±0.20 m-sec–1. The deviations increase with increasing ionic strength (±0.08 m-sec–1 at I=0.1 and ±0.25 m-sec–1 at I=0.5 m).The sound speed of seawater have also been estimated from 0 to 40°C, 0.1 to 0.7 ionic strength and 0 to 1000 bars. The estimates were found to be in good agreement (±0.4 m-sec–1) with the measured values.These results indicate that reasonable estimates of the adiabatic PVT properties of dilute mixtures of electrolyte solutions can be made using the additivity principle, without excess mixing terms.  相似文献   

8.
The speed of sound of mixtures of the six possible combinations of the major sea salt ions (Na+, Mg2+, Cl, and SO 4 2– ) have been determined at I=3.0 and at 25°C. The results have been used to determine the changes in the adiabatic compressibility of mixing Km the major sea salts. The values of Km have been fit to the equation Km=y2y3I2[k0+k1(1-2y3)] where yi is the ionic strength fraction of solute i, k0 and k1 are parameters related to the interactions of like-charged ions. The Young cross-square rule is obeyed to within ±0.04×10–6 cm3-kg–1-bar–1. A linear correlation was found between the compressibility k0 and volume v0 interaction parameters (104k0=–0.24+3.999 v0, s=0.15) in agreement with out earlier findings. Estimates of the sound speeds for the cross square mixtures (NaCl+MgSO4 and MgCl2+Na2SO4) were made using the equations of Reilly and Wood. The estimated sound speeds were found to agree on the average with the measured values to ±0.36 m-sec–1.  相似文献   

9.
Dissociative and nondissociative electron attachment in the electron impact energy range 0–14 eV are reported for SOF2 SOF4, SO2F2, SF4, SO2, and SiF4 compounds which can be formed by electrical discharges in SF6. The electron energy dependences of the mass-identified negative ions were determined in a time-of-flight mass spectrometer. The ions studied include F and SOF 2 –* from SOF2; SOF 3 and F from SOF4; SO2F 2 –* , SO2F, F 2 , and F from SO2F2; SF 4 –* and F from SF4; O, SO, and S from SO2; and SiF 3 and F from SiF4. Thermochemical data have been determined from the threshold energies of some of the fragment negative ions. Lifetimes of the anions SOF 2 –* , SO2F 2 –* , and SF 4 –* are also reported.  相似文献   

10.
Infrared spectra in digitized form were measured for NaNO3 and [Na·C221]+NO 3 solutions in DMSO-d6 between 1150 and 1500 cm–1 using a technique and instrumentation that obtains each point of the average absorbance spectrum at the same (reduced) noise level. Similar spectra were also obtained for the solvent and the Na+ complexed cryptand C221 and used to remove the contribution of these entities from the above spectra. By taking appropriate differences of spectra, it was possible to reveal both bands of the contact ion pair in the NaNO3/DMSO-d6 solution-removing one from under the strong band of the D3h site—and to show the presence of three ion sites in this solution. The third site is tentatively identified as a close ion pair. Two ion sites are also identified in the [Na·C221]+NO 3 /DMSO-d6 solution.Paper X in the series, Studies of Solution Character, by Molecular Spectroscopy.  相似文献   

11.
Raman spectra of aqueous Zn(II)–perchlorate solutions were measured over broad concentration (0.50–3.54 mol-L–1) and temperature (25–120°C) ranges. The weak polarized band at 390 cm–1 and two depolarized modes at 270 and 214 cm–1 have been assigned to 1(a 1g), 2(e g), and 5(f 2g) of the zinc–hexaaqua ion. The infrared-active mode at 365 cm–1 has been assigned to 3(f 1u). The vibrational analysis of the species [Zn(OH2) 2 + ] was done on the basis of O h symmetry (OH2 as point mass). The polarized mode 1(a 1g)-ZnO6 has been followed over the full temperature range and band parameters (band maximum, full width at half height, and intensity) have been examined. The position of the 1(a 1g)-ZnO6 mode shifts only about 4 cm–1 to lower frequencies and broadens by about 32 cm–1 for a 95°C temperature increase. The Raman spectroscopic data suggest that the hexaaqua–Zn(II) ion is thermodynamically stable in perchlorate solution over the temperature and concentration range measured. These findings are in contrast to ZnSO4 solutions, recently measured by one of us, where sulfate replaces a water molecule of the first hydration sphere. Ab initio geometry optimizations and frequency calculations of [Zn(OH2) 2 + ] were carried out at the Hartree–Fock and second-order Møller–Plesset levels of theory, using various basis sets up to 6-31 + G*. The global minimum structure of the hexaaqua–Zn(II) species corresponds with symmetry T h. The unscaled vibrational frequencies of the [Zn(OH2) 2 + ] are reported. The unscaled vibrational frequencies of the ZnO6, unit are lower than the experimental frequencies (ca. 15%), but scaling the frequencies reproduces the measured frequencies. The theoretical binding enthalpy for [Zn(OH2) 2 + ] was calculated and accounts for ca. 66% of the experimental single-ion hydration enthalpy for Zn(II).Ab initio geometry optimizations and frequency calculations are also reported for a [Zn(OH2) 2 18 ] (Zn[6 + 12]) cluster with 6 water molecules in the first sphere and 12 in the second sphere. The global minimum corresponds with T symmetry. Calculated frequencies of the zinc [6 + 12] cluster correspond well with the observed frequencies in solution. The 1-ZnO6 (unscaled) mode occurs at 388 cm–1 almost in perfect correspondence to the experimental value. The theoretical binding enthalpy for [Zn(OH2) 2 18 ] was calculated and is very close to the experimental single ion-hydration enthalpy for Zn(II). The water molecules of the first sphere form strong hydrogen bonds with water molecules in the second hydration shell because of the strong polarizing effect of the Zn(II) ion. The importance of the second hydration sphere is discussed.  相似文献   

12.
Equations in the ion-interaction (Pitzer) system are derived for the volume change on mixing any combination of the sea salts NaCl, Na2SO4, MgSO4, MgCl2 at constant ionic strenth. For these mixings of different charge types, the equations include complex differences of pure electrolyte terms. Recently measured data for each of the pure electrolytes provide these pure electrolyte terms. Other recent measurements on the volume change on mixing are compared with values calculated from the equations. At 25°C there is no need to introduce the mixing terms based on differences in the interactions of ions of the same sign. At other temperatures, the agreement without the mixing terms is good, but significant improvement is obtained by inclusion of the binary mixing terms Cl,SO 4 v and Na,Mg v . The equations and parameters can then predict the volumetric properties of any mixed solution of these salts over the range 0–100°C and to at least 3 mol-kg–1 ionic strength.  相似文献   

13.
Viscosities for aqueous NH4Cl and tracer diffusion coefficients for22Na+,36Cl, HTO, and CH3OH, acetone and dimethylformamide (all14C-labelled) in NH4Cl supporting electrolyte are reported for 25°, together with tracer diffusion coefficients for22Na+,36Cl, and14CH3OH in 1M KI, and for14CH3OH in 1M MgCl2. The diffusion coefficient of HTO in NH4Cl solutions is slightly larger, for most of the concentration range investigated (0.05 to 4.5 M), than the value for pure water and is almost unaffected by the supporting electrolyte up to about 4M. Similar behavior is shown by CH3OH, acetone and dimethylformamide in NH4Cl solutions. Onsager limiting law behavior is approached by Cl at NH4Cl concentrations in the 0.05–0.1M region but at much lower concentrations by Na+.  相似文献   

14.
205Tl and1H magnetic resonance frequencies have been determined for liquid ammonia solutions of TlClO4 and TlNO3 as a function of electrolyte concentration (5×10–4 to 9.8 M) and temperature. The dependence of the resonance frequency on concentration suggests the presence of free, fully solvated thallium ions, ion pairs, and higher-order ion aggregates. Analysis of lowconcentration205Tl data between 0 and 30°C allowed the determination of TlNO3 ion-pair association constants and thermodynamic parameters (HA=+6.5 kcal-mole–1, SA=+36 e.u.). A preciptous decrease in205Tl resonance frequency was observed for NH3 to TlNO3 mole ratios below 3:1, suggesting the formula (NH3)3Tl+ NO 3 for the fully solvated, contact ion pair.  相似文献   

15.
The oxidation of Fe(II) with H2O2 has been measured in NaCl and NaClO4 solutions as a function of pH, temperature T (K) and ionic strength (M, mol-L–1). The rate constants, k (M–1-sec–1), d[Fe(II)]/DT=-k[Fe(II)][2O2]at pH=6.5 have been fitted to equations of the formlog k = log k0+ AI 1/2+BI+CI 1/2/T Where log k0=15.53-3425/T in water; A=–2.3, –1.35; B=0.334, 0.180; and C=391, 235, respectively, for NaCl (=0.09) and NaClO4 ( =0.08). Measurements made in NaCl solutions with added anions yield rates in the order B(OH) 4 >HCO 3 >ClO 4 >Cl>NO 3 >SO 4 2– and are attributed to the relative strength of the interactions of Fe2+ or FeOH+ with these anions. The FeB(OH) 4 + species is more reactive while the FeCO 3 0 , FeCl+, FeNO 3 + and FeSO 4 0 species are less reactive than the FeOH+ ion pair. The general trend is similar to our earlier studies of the oxidation of Fe(II) with O2 except for B(OH) 4 . The effect of pH on the logk was found to be a quadratic function of the concentration of H+ or OH from pH=4 to 8. These results have been attributed to the different rate constants for Fe2+ (k0) and FeOH+ (k1) which are related to the measured k by, k=k0Fe + k1FeOH, where i is the molar fraction of species i. The rates increase due to the greater reactivity of FeOH+ compared to Fe2+. k0 is independent of composition and ionic strength but k1 is a function of ionic strength and composition due to the interactions of FeOH+ with various anions.  相似文献   

16.
A pre-resonance Raman study of the yellow -quinol/SO2 clathrate has been carried out using 609.8, 586.8, 514.5, 488.0 and 457.9 run excitation. Pre-resonance enhancement is observed for the guest vl (Al) band at 1147 cm–1 and the host band at 1257 cm–1. These observations are consistent with a charge transfer interaction arising from the LUMO of S02 (S 3pz) and the HOMO of quinol, which consists mainly of the ring electrons.  相似文献   

17.
Aqueous solutions of sodium chloride, potassium chloride, sodium sulfate, and potassium sulfate can be mixed in six ways to give ternary mixtures. Two of these have already been studied and results are now presented for the remaining four systems: H2O–NaCl–K2SO4, H2O–Na2SO4–K2SO4, H2O–KCl–Na2SO4, and H2O–KCl–K2SO4.  相似文献   

18.
A pronounced effect of structural heterogeneity (cracks) of glass-like solutions of 4.9M H2SO4 on their radiothermoluminescence (RTL) was found. In perfect glasses one RTL peak was observed at 115 K. Additional luminescence peaks appeared at 165, 195, and 240 K in glasses having cracks. The effect observed was explained by elevated thermal stability of the SO4 radical stabilized on the surface of sulfuric acid crystal hydrates: H2SO4 · 4H2O and H2SO4· 6.5H2O.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 6, pp. 1566–1568, June, 1996.  相似文献   

19.
Addition of water to stoichiometric 100% sulfuric acid increases the density untila maximum results near 87 mole% H2SO4. The density and conductivity maximaand viscosity minimum, the latter two near 75 mole%, are direct macroscopicresponses to microscopic quantum mechanical properties of H3O+ and of nearlysymmetric H-bond double-well potentials, as follows: (1) lack of H bonding tothe O atom of H3O+; (2) short, 2.4–2.6 A, O—O distances of nearly symmetricH bonds; and, (3) increased mobility of protons in such short H bonds, give riseto the density maximum via (1) and (2); (1) produces the viscosity minimum;and the conductivity maximum results from (2) and (3). A pronounced minimumnear 1030 cm–1 in the symmetric SO3 stretching Raman frequency of HSO4 ,observed near 45 mole% also results from double-well effects involving the shortH bonds of direct hydronium ion—bisulfate ion pair interactions. Estimates of theconcentrations of the (H3O+)(HSO4 ) and (H2SO4)(HSO4 ) pair interactions weredetermined from Raman intensity data and are given for compositions between42–100 mole%  相似文献   

20.
The solubility of CaSO3·1/2H2O(c) was studied under alkaline conditions (pH>8.2), in deaerated and deoxygenated Na2SO3 solutions ranging in concentration from 0.0002 to 0.4M and in CaCl2 solutions ranging in concentration from 0.0002 to 0.01M, for equilibration periods ranging from 1 to 7 days. Equilibrium was approached from both the over- and the under-saturation directions. In all cases, equilibrium was reached in <1 days. The aqueous Ca2+–SO 3 2– ion interactions can be satisfactorily modeled using either ion-association or ion-interaction aqueous thermodynamic models. In the ion-association model, the log K°=2.62±0.07 for Ca2++SO 3 2– CaSO 3 0 . In the Pitzer ion-interaction model, the binary parameters (0) and (1) for Ca2+–SO 4 2– were used, and the value of (2) was determined from the experimental data. As expected given the strong association constant, the value of (0) was quite small (about –134). We feel a combination of the two models is most useful. The logarithm of the thermodynamic equilibrium constant (K°) of the CaSO3·1/2H2O(c) solubility reaction (CaSO3·1/2H2O(c)Ca2++SO 3 2+ +0.5H2O) was found to be –6.64±0.07.  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