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.
Ion diffusion kinetics has been studied using the data of conductivity measurements for aqueous solutions of sodium selenite with different concentrations and at different temperatures. Molecular and ionic self-diffusion coefficients have been determined for infinitely dilute solutions in the temperature range 288 K-313 K. The limiting values of ion mobility and changes in the energies of translation of water molecules from ions’ hydration shell have been found. At elevated temperatures, ΔE tr 0 increases for both ions in direct proportion to the crystallographic radius of the latter. Ion hydration numbers at 298 K have been calculated. The results of this study are interpreted in the light of Samoilov’s theory on positive and negative hydration of ions.Original Russian Text Copyright © 2004 by L. T. Vlaev and S. D. Genieva__________Translated from Zhurnal Strukturnoi Khimii, Vol. 45, No. 5, pp. 870–876, September–October, 2004.  相似文献   

2.
Vanillin–Schiff’s bases (VSB) were examined as thermal stabilizers and co-stabilizers for rigid poly(vinyl chloride) (PVC) in air at 180 °C. Their high stabilizing efficiency were shown by their high thermal stability value (Ts), which is the time elapsed for the detection of HCl gas, if compared with dibasic lead carbonate and cadmium–zinc soap reference stabilizers used industrially, with better extent of discoloration. Blending these derivatives with reference stabilizers in different ratios greatly lengthens the thermal stability and the extent of discoloration of the PVC.Condensation products of Vanillin with amines are very active biologically, besides having good complexation ability with metal ions. The Ni2+ and Co2+ complexes of VSB derivatives gave better thermal stability and less discoloration than the parent organic stabilizer. Also, blending these complexes with either of the used reference stabilizers in different ratios gave better thermal stability and lower extent of discoloration. Thermogravimetric analysis confirmed the improved stability of PVC in the presence of the VSB derivatives, compared to blank PVC, PVC stabilized with reference stabilizers and PVC stabilized with binary mixture of VSB derivatives with reference stabilizer.The stabilizing efficiency of Vanillin–Schiff’s base (VSB) derivatives is attributed to the replacement of the labile chlorine atoms on the PVC chains by a relatively more stable moiety of the organic stabilizer.  相似文献   

3.
We report molecular dynamics studies on the effect of CCD (chlorinated cobalt-dicarbollide) anions on the Eu3+ lanthanide cation extraction by a calix[4]arene-CMPO ligand L, focusing on the water–‘oil’ interface, where ‘oil’ is modelled by chloroform. The free L ligand and its EuL3+ complex are found to adsorb and to concentrate at the interface, but are too hydrophilic to be extracted. Addition of CCD anions in diluted conditions (either covalent linked to L or as separated CCD H3O+ ions) also leads to adsorption of these species at the interface. However, at high concentrations, CCD anions saturate the interface and promote the extraction of EuL3+ to the oil phase. Another important feature concerns the uncomplexed Eu(CCD)3 salt: accumulation of CCD anions at the interface creates a negative potential which attracts the hydrated Eu3+ ions, therefore facilitating their complexation by interfacial ligands. These features allow us to better understand the synergistic effect of lipophilic anions in the assisted liquid-liquid extraction of trivalent M3+ lanthanide or actinide cations. To cite this article: B. Coupez, G. Wipf, C. R. Chimie 7 (2004).

Résumé

Synergie due aux anions dicarbollides lors de l’extraction d’ions lanthanides M3+ par des calix[4]arènes : simulations de dynamique moléculaire à l’interface eau–« huile ». Nous étudions par simulations de dynamique moléculaire l’effet de synergie dû aux anions CCD (cobalt-dicarbollides) lors de l’extraction de Eu3+ par un calix [4]arène L, en se focalisant sur l’interface eau–« huile », l’huile étant modélisée par du chloroforme. On montre que le ligand L et son complexe EuL3+ s’adsorbent à l’interface, mais sont trop hydrophiles pour être extraits. L’addition d’anions CCD (qu’ils soient sous la forme d’ions CCD H3O+ séparés ou greffés de façon covalente au calixarène) conduit aussi à l’adsorption de ces espèces à l’interface. Cependant, aux plus fortes concentrations, les anions CCD saturent l’interface et induisent l’extraction du complexe EuL3+ vers l’huile. Un autre résultat remarquable concerne les sels Eu(CCD)3 : l’accumulation des anions CCD à l’interface y crée un potentiel négatif, ce qui attire les cations Eu3+ et facilite ainsi leur complexation par des ligands à l’interface. Ces résultats permettent de mieux comprendre l’effet de synergie dû aux anions CCD lors de l’extraction d’ions lanthanides ou actinides M3+ et, d’une manière générale, ce qui se passe à l’interface entre l’eau et des liquides non miscibles. Pour citer cet article : B. Coupez, G. Wipf, C. R. Chimie 7 (2004).  相似文献   

4.
Mechanical activation (MA) of the LiOH+V2O5 and Li2CO3+V2O5 mixtures followed by brief heating at 673 K was used to prepare dispersed Li1+xV3O8. It was shown that structural transformations during MA are accompanied by reduction processes. EPR spectra of Li1+xV3O8 are attributed to vanadyl VO2+ ions with weak exchange interaction. The interaction of localized electrons (V4+ ions) with electron gas (delocalized electrons), which is exhibited through the dependence of EPR line width of vanadium ions versus measurement temperature (C–S–C relaxation), is revealed. It is shown that C–S–C relaxation is different for intermediate and final products. The properties of mechanochemically prepared Li1+xV3O8 are compared with those of HT-Li1+xV3O8, obtained by conventional solid state reaction. Mechanochemically prepared Li1+xV3O8 is characterized by a similar amount of vanadium ions, producing electron gas, but a higher specific surface area.  相似文献   

5.
The densities of mixtures of the six possible combinatons of the major sea salts (NaCl, Na2SO4, MgSO4, and MgCl2) were determined at constant ionic strengths of I=1.0 and I=3.0 at 25°C. The results are used to determine the volume changes for mixing (V m ) the major sea salts. The values of V m were fit to equations of the form V m where y i is the molal ionic strength fraction of solute i, and 0 and 1 are parameters related to the interaction of like-charged ions. The cross-square rule was found to hold at both ionic strengths. Density estimates were made without and with the addition of volume of mixing terms to Young's Rule and compared to the experimental values. The densities calculated with the addition of volume of mixing terms gave better estimates, demonstrating that the densities of concentrated brines can be more accurately estimated using V m terms. The equations of Reilly and Wood which include the cross-square rule were used to estimate the densities of the cross mixtures (NaCl–MgSO4 and MgCl2–Na2SO4). The estimated densities agree with the measured values to within ±30 ppm at I=1.0 and ±125 ppm at I=3.0.  相似文献   

6.
Electrospray‐generated precursor ions usually follow the ‘even‐electron rule’ and yield ‘closed shell’ fragment ions. We characterize an exception to the ‘even‐electron rule.’ In negative ion electrospray mass spectrometry (ES‐MS), 2‐(ethoxymethoxy)‐3‐hydroxyphenol (2‐hydroxyl protected pyrogallol) easily formed a deprotonated molecular ion (M‐H)? at m/z 183. Upon low‐energy collision induced decomposition (CID), the m/z 183 precursor yielded a radical ion at m/z 124 as the base peak. The radical anion at m/z 124 was still the major fragment at all tested collision energies between 0 and 50 eV (Elab). Supported by computational studies, the appearance of the radical anion at m/z 124 as the major product ion can be attributed to the combination of a low reverse activation barrier and resonance stabilization of the product ions. Furthermore, our data lead to the proposal of a novel alternative radical formation pathway in the protection group removal of pyrogallol. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

7.
Densities (ρ)of the binary systems of {difurylmethane + (ethanol or propan-1-ol or butan-1-ol or pentan-1-ol or hexan-1-ol)} have been measured with an Anton Paar DMA 4500 vibrating-tube densimeter over the entire composition range at 298.15,K and atmospheric pressure. Excess molar volumes (V m E ) of each binary system were determined and correlated by the Redlich-Kister equation. Limiting (V i E,∞) and excess partial molar volumes (V i E ) of components of each binary system have been calculated to provide insight into the intermolecular interactions present and the packing efficiencies. The results have been discussed in terms of specific intermolecular interactions, dispersive forces and structural effects.  相似文献   

8.
The densities of tetraalkylammonium bromide, R4NBr (R = Et, Pr, Bu, Hex, Hep, Oct), solutions in dimethylformamide have been measured for the composition range (0.05–0.4) mol-kg−1 at 25 C. Apparent molar Vφ and limiting partial molar volumes 2o of the electrolytes have been evaluated. Using the extrapolation values, the limiting partial molar volumes of the tetraalkymammonium ions (io) have been calculated. Analysis of different contributions to the ionic io indicated partial penetration of solvent molecules into the van der Waal’s volume of tetraalkylammonium (TAA) ions.  相似文献   

9.
Michael H. Palmer   《Chemical physics》2009,360(1-3):150-161
The 1,2,5-oxadiazole VUV absorption spectrum in the range 5–11.5 eV, shows broad bands centred near 6.2, 7.1, 8.3, 8.8, 10.6 and 11.3 eV. Rydberg states associated with three ionisation energies (IE) were identified in the complex fine structure above 8.7 eV. Electronic vertical excitation energies for singlet and triplet valence, and Rydberg states were computed using ab initio multi-reference multi-root CI methods. There is generally a good correlation between the envelope of the theoretical intensities and the experimental spectrum. The nature of the more intense calculated Rydberg states, and positions of the main valence and Rydberg bands are discussed. The lowest triplet, singlet and Rydberg 3s excited states have equilibrium structures that are non-planar with CS symmetry, in a chair-like orientation where the O and H atoms lie out of the NCCN plane. This finding is consistent with the doubling of the low energy UV spectral lines [B.J. Forrest, A.W. Richardson, Can. J. Chem., 50 (1972) 2088].The nearly degenerate IE of the UV-photoelectron spectrum (UV–PES, Palmer et al. 1977) makes analysis of the VUV spectrum difficult, leading to the necessity for reinvestigation. Vertical studies (IEV) using CI, Tamm–Dancoff (TDA) and Green’s Function (GF) methods all gave similar results, with near degeneracy of the first 3IEV confirming the earlier study. Studies of the adiabatic IE (IEA) using CCSD(T) and B3LYP methods, showed the energy sequence 2A2 < 2B1 < 2B2, but these states are all saddle points, in contrast to the 4th state (2A1) which is a minimum. In contrast, MP2 study of the 2B2 state showed a minimum, with only two saddle points.Complete minima were found after minor twisting of the structures. The lowest energy cationic state is 2A (CS), which closely resembles the 2B2 state. The O–N–C–C skeleton is twisted by 8°. The corresponding 2A state (CS) is effectively identical to the 2B1 state. Attempts to find minima for other symmetry states were unsuccessful.  相似文献   

10.
Summary From kinetic studies of18O-exchange between aqueous solutions of K12V18O42 · 16H2O and water it is concluded that [V18O42]12– exists as a discrete ion in this medium. The rate of exchange is relatively slow (t1/2 ca. 6×104s at 0°C) and obeys the McKay equation within the precision obtainable. The sensitivity of the ion to its environment in the solid state, leading to induced exchange, prevented a decision as to whether all oxygens are kinetically equivalent. It is clear, however, that the ions remains intact for long periods in solution in the absence of air.  相似文献   

11.
Excess molar volumes, V E m, at 25°C and atmospheric pressure, over the entire composition range for binary mixtures of methanol, ethanol, 1-propanol, 1-butanol, 1-pentanol, 1-hexanol, 1-heptanol, and 1-octanol with-methylbutylamine are reported. They are calculated from densities measured with a vibrating-tube densimeter. All the excess volumes are large and negative over the entire composition range. This indicates strong interactions between unlike molecules, which are greatest for the system involving methanol, characterized by the most negative V E m. For the other solutions, V E m at equimolar composition, is approximately the same. The V E m curves vs. mole fraction are nearly symmetrical. The ERAS model is applied to 1-alkanol + N-methylbutylamine, and 1-alkanol + diethylamine systems. The ERAS parameters confirm that the strongest interactions between unlike molecules are encountered in solutions with methanol. The model consistently describes V E m and excess molar enthalpies H E m of the mixtures studied.  相似文献   

12.
The infrared, Raman, and inelastic neutron scattering (INS) spectra of two ortho-hydroxy aryl Schiff’s bases, 2-(N-methyliminoethyl)-phenol and 2-(N-methyl-α-iminoethyl)-phenol, were recorded. Ab initio molecular orbital calculations employing the DFT (B3LYP) method with the 6-31G** basis set for both compounds were done. Assignments of vibrational modes within the 3500–50 cm−1 spectral region were carried out. On the basis of the DFT calculations, four rotomers of 2-(N-methyl-α-iminoethyl)-phenol were analysed.  相似文献   

13.
The interfacial bending moment and curvature elastic moduli are related theoretically to the surface potential, ΔV, which is liable to direct measurement. The dependence of the interfacial bending properties on the position of the Gibbs dividing surface is investigated. As an application, experimental data for the surface potential of micellar surfactant solutions containing Al3+ions are analyzed. The changes in the bending moment, as determined from ΔVpotential, correlate with the transition from spherical to cylindrical micelles induced by the Al3+ions. The results can be important for interpretation of data for formation of microemulsions, flocculation in emulsions, fluctuation capillary waves at interfaces and biomembranes, interactions between inclusions in lipid bilayers, etc.  相似文献   

14.
15.
Cu2+ binding on γ-Al2O3 is modulated by common electrolyte ions such as Mg2+, , and in a complex manner: (a) At high concentrations of electrolyte ions, Cu2+ uptake by γ-Al2O3 is inhibited. This is partially due to bulk ionic strength effects and, mostly, due to direct competition between Mg2+ and Cu2+ ions for the SO surface sites of γ-Al2O3. (b) At low concentrations of electrolyte ions, Cu2+ uptake by γ-Al2O3 can be enhanced. This is due to synergistic coadsorption of Cu2+ and electrolyte anions, and . This results in the formation of ternary surface species (SOH2SO4Cu)+, (SOH2PO4Cu), and (SOH2HPO4Cu)+ which enhance Cu2+ uptake at pH < 6. The effect of phosphate ions may be particularly strong resulting in a 100% Cu uptake by the oxide surface. (c) EPR spectroscopy shows that at pH  pHPZC, Cu2+ coordinates to one SO group. Phosphate anions form stronger, binary or ternary, surface species than sulfate anions. At pH  pHPZC Cu2+ may coordinate to two SO groups. At pH  pHPZC electrolyte ions and are bridging one O-atom from the γ-Al2O3 surface and one Cu2+ ion forming ternary [γ-Al2O3/elecrolyte/Cu2+] species.  相似文献   

16.
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   

17.
The critical micellar concentration of sodium dodecyl sulfate is strongly altered bytris(hydroxy-methyl)methylammonium ions. The effect of buffer solutions containing this weak electrolyte as the counterion source has been studied using various concentrations of the acid–base system as well as modifying the pH. Results show that counterion concentrations ranging from 0 to 340 × 10−3M induce an appreciable diminution of the critical micellar concentration from 8 to 0.7 × 10−3M. The analysis of data suggests that the critical micellar concentration of sodium dodecyl sulfate depends on the concentration of weak electrolytes in a way very similar to that of strong electrolytes.  相似文献   

18.
The electronic spectra of solid iron(III) vanadates FeVO4 and Fe2V4O13 were investigated by the diffuse reflectance technique in the spectral range 12 500–50 000 cm−1. The spectra of investigated vanadates contain 2–3 intensive CT bands in the UV region and two lowest energy dd bands in the 12 000–22 000 cm−1 range. The presence of the weak bands for FeVO4 and Fe2V4O13 at 16 500 cm−1 and 20 500 cm−1 points to the lattice deffects (oxygen deficiency and the presence of the V4+ ions) in the structure of investigated vanadates.  相似文献   

19.
The binary mixtures of 7 hexoses and 20 amino acids were investigated by electrospray ionization ion trap mass spectrometry (ESI‐ITMS). The adduct ions of the amino acid and the hexose were detected for 12 amino acids but not for the other 8 amino acids which are basic acidic amino acids and amides. The ions of amino acid–hexose complexes were further investigated by tandem mass spectrometry (MS/MS), and some of them just split easily into two parts whereas the others gave rich fragmentation, such as the complex ions of isoleucine, phenylalanie, tyrosine, and valine. We found that hexoses could be complexed by two molecules of valine but only by one molecule of the other amino acids. Among seven kinds of valine–hexose complexes coordinated by potassium ion, the MS2 spectra of the ion at m/z 453 yielded unambiguous differentiation. And the fragmentation ions are sensitive to the stereochemical differences at the carbon‐4 of hexoses in the complexes, as proved by the MS2. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

20.
The thiomannoside is a key intermediate in the synthesis of polymannosaccharide. But its crystallographic study has not been reported. X-ray crystal structure analysis shows that the crystal of phenyl 2,3,4,6-tetra-O-acetyl-1-thio-α-D-mannopyranoside is in orthorhombic system, has P212121 space group. Its unit cell dimensions are a=9.1731(11), b=11.574(2), and c=21.199(3) Å, α=β=γ=90.00°, Z=4, V=2250.6(6) Å3, Dc=1.300 g/cm3. The crystallographic study reveals that the thiomannoside is an α-anomer. Most interestingly, the structure provides an evidence of existing C–HO and C–HS intramolecular and intermolecular hydrogen bonds.  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