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.
The dielectric properties of mixed monolayers of per-(6-amino-2,3-di-O-hexyl) β-CD hydrochloride (NH3-β-CD-OC6) and 1,2 dipalmitoyl, 3-sn-phosphatidic acid (DPPA) have been assessed using surface potential measurements at constant area. From the comparison of these surface potential (ΔV) versus surface density (δ) relationships with those of surface pressure (π) against surface density (δ) it was apparent that the increase in the NH3-β-CD-OC6 content in mixed films gave rise to a gradual increase in the saturation value of the surface potential (ΔVmax). This potential for pure DPPA was found to be equal to 396 mV and for pure CD 554 mV. The ΔVmaxvalues reflect the onset of reorientation effects that arrive at molar areas before the collapse of these films. Independently of reorientation effects, the obtained results strongly indicate that the dipolar term contributing to the overall ΔVvalue was for NH3-β-CD-OC6 due to the hydration of its NH+3group. For both DPPA and NH3-β-CD-OC6 molecules the contribution of the electric double layer (Ψ) was calculated and was found for DPPA and NH3-β-CD-OC6 to be equal to −249 and +252 mV, respectively. These calculated Ψ values made possible the evaluation of dipole moments for NH3-β-CD-OC6 and DPPA monolayers which revealed a marked difference in dipolar properties between these two film forming components. In contrast to DPPA which exhibited a decrease in the surface dipole moment (μ) with the decrease inA, NH3-β-CD-OC6 displayed an increase in μwith the decrease inAforAvalues above 580 Å2. Below this value μdecreases with decreasing molecular area and this variation arises from a change in the polarity of the electric double layer arising from interactions with the complementary anion. The differences in dielectric properties between the two film forming molecules have been attributed to modification, during compression, in the structure of the interfacial water bound to the cyclodextrin.  相似文献   

2.
Adsorption of ammonia and nitrogen oxide on V2O5/Al2O3 samples of different degrees of reduction has been studied by IR spectroscopy. On an oxidized surface, ammonia is coordinated by V3+ and V4+ ions to form ammonium ions; NO is not adsorbed. On a reduced surface, the coordination of NO by V3+, V4+, and V5+ ions is observed, which results in the formation of nitrosyl complexes. A strong mutual influence between NO and NH3 occurs during coadsorption or consecutive interaction on a reduced catalyst surface.Translated fromIzvestiya Akademii Nauk. Seriya Khimicheskaya, No. 2, pp. 236–241, February, 1994.  相似文献   

3.
Surface properties (Langmuir monolayer) of two different cerebrosides which are extracted from the sea cucumber (Bohadschia argus) were investigated. A main difference in chemical structure of cerebroside between BAC-2a and BAC-4 is their head groups (glucose and galactose, respectively). Furthermore, miscibility and interaction between dipalmitoylphosphatidylcholine (DPPC) and cerebrosides (BAC-2a and BAC-4) in the monolayer have been systematically examined. The surface pressure (π)−area (A), the surface potential (ΔV)−A, and the dipole moment (μ)−A isotherms for monolayers of DPPC, cerebrosides, and their binary combinations have been measured using the Wilhelmy method and the ionizing electrode method. BAC-4 forms a stable liquid-expanded (LE) monolayer, whereas BAC-2a has a first-order phase transition from the LE phase to the liquid-condensed (LC) state on 0.15 M NaCl at 298.2 K. The fundamental properties for each cerebroside monolayer were elucidated in terms of the surface dipole moment based on the three-layer model [R.J. Demchak, T. Fort Jr., J. Colloid Interface Sci. 46 (1974) 191–202] for both cerebrosides and the apparent molar quantity change (Δsγ, Δhγ, and Δuγ) for BAC-2a. In addition, their miscibility with DPPC was examined by the variation of the molecular areas and the surface potentials as a function of cerebroside mole fractions, the additivity rule. The miscibility was also confirmed by constructing the two-dimensional phase diagrams. The phase diagrams for the both binary systems were of negative azeotropic type. That is, the two-component DPPC/BAC-2a and DPPC/BAC-4 monolayers are miscible. Furthermore, the Joos equation for the analysis of the collapse pressure of binary monolayers allowed calculation of the interaction parameter and the interaction energy between the DPPC and cerebroside monolayers. The miscibility in the monolayer state was also confirmed by the morphological observation with Brewster angle microscopy (BAM), fluorescence microscopy (FM), and atomic force microscopy (AFM).  相似文献   

4.
The formation of metal complexes between water-soluble polymers, poly(vinyl alcohol) [PVA], poly(N-vinylpyrrolidone) [PVP], poly(acrylamide) [PAAm] and poly(ethylene oxide) [PEO] with trivalent metal ions, Fe3+, Cr3+, and V3+ were studied by using differential pulse polarography (DPP). The general experimental observation is the shift of totally reversible reduction peaks (M3++Hg+eM2++Hg) towards more negative potentials when the complexing water-soluble polymers are added to the solution of trivalent metal ions. The negative shift in potential permitted the determination of complex formation constants (Kf) between trivalent metal ions and water soluble polymers. The complex formation constants for Fe3+, Cr3+, and V3+ ions with these polymers increased in the order of V3+>Cr3+>Fe3+.  相似文献   

5.
Excess volumes of mixing for six possible binary combinations of solutions of NaCl, KCl, NaBr and KBr have been determined at constant ionic strengths of 1.000 and 2.000 mol-kg− 1 at 298.15 K using a dilatometer in the water + 1,4-dioxane mixed-solvent system. Pitzer’s ion-interaction model has been utilized to obtain binary and triplet interaction parameters, i.e., θV and ψV. The data were also analyzed by the Friedman Model and it is suggested that interactions between solvated ions are dictated not only by coulombic interactions but also by appreciable asymmetric effects. The data are dependent on the nature of the common ion and do not support Young’s cross-square rule. The deviation from the cross-square rule is considered to arise from preferential solvation of the ions and ion clusters in the mixed-solvent system as reflected by the appreciable contribution of triplet interactions.  相似文献   

6.
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   

7.
The potential surface for the reaction H2CO+H → HCO+ + H2 has been studied by ab initio SCF calculations, using gaussian-type basis functions. A saddle point on the surface has been found, and a reaction path is proposed to explain the observed release of kinetic energy. The energy of activation and ΔE for the reaction have been estimated.  相似文献   

8.
A temperature study was performed on micelle formation of a series of homologous cationic surfactants having organic counterions (alkanesulfonates) with carbon numbers ranging from 1 to 4: dodecylammonium salts of methanesulfonate (DAMS), ethanesulfonate (DAES), propanesulfonate (DAPS), and butanesulfonate (DABS) in water. The critical micelle concentrations (CMCs) and the degree of counterion binding (β) were determined at different temperatures ranging from 5 to 50°C by means of conventional electric conductance measurements. From the temperature dependence of β as well as CMC, Gibbs energy ΔG0m, enthalpy ΔH0m, and entropy ΔS0m, on micelle formation, were estimated for the respective surfactants. As for the temperature dependence of CMC for these surfactants, the temperature-CMC curves have a minimum around 30°C and show that the CMC at each temperature is lowered by about 3 mmol dm-3 per methylene group in the alkyl chain of the counterions. The relationship between β and temperature suggested that the counterion of MS- behaves most similarly to common univalent ions such as halide ions. In contrast, PS- and BS-, having a stronger ability to lower CMC and to promote association of surfactant ions with counterions as well as of surfactant ions themselves, behave more like those of surfactant ions, and ES- shows the most complicated character between those of common univalent ions and organic ions. However, the temperature dependence of enthalpy change, ΔH0m demonstrates that these four surfactants are divided into two groups: (1) DAMS and DAES and (2) DAPS and DABS. In addition, the entropy change ΔS0m as a function of alkyl chain length gives evidence that the contribution of the entropy term to the Gibbs energy on micelle formation clearly separates between DAES (m = 2) and DAPS (m = 3). A similar discontinuity is found even in the plot of ΔG0m versus carbon atom number of alkyl chain, m, and in the plot of ΔG0m versus estimated hydrodynamic radius of counterions. All the results obtained have indicated that lengthening the alkyl chains initially hinders micelle formation, but the longer chains are markedly effective in lowering the CMC and probably in increasing the aggregation number, owing to enhanced hydrophobic interaction between counterion and the micellar surface and/or core.  相似文献   

9.
Mössbauer spectra of the Fe1+xV2−xO4 spinel solid solutions are taken to investigate the cation distribution. Room temperature spectra can be interpreted by assuming that the cation distribution is represented approximately as Fe2+[Fe3+xV3+2−x]O4 for 0 x 0.35 and Fe3+[Fe2+Fe3+x−1V3+2−x]O4 for 1 x 2 and the ionic valence arrangement changes from the 2-3-3 type (Fe2+[Fe3+xV3+2−x]O4) to the 3-2-3 one (Fe3+[Fe2+V3+]O4) in the range 0.35 x 1. Fe2VO4 is found to be 3-2-3 spinel, Fe3+[Fe2+V3+]O4. Its paramagnetic spectrum at 473°K is, however, composed of a broad single line with isomer shift value of 0.61 mm/sec relative to stainless steel, in which the line splitting due to the ferric and ferrous ions is rendered indistinguishable.  相似文献   

10.
The kinetics of the acid dissociation of copper(II) complexes of novel C-functionalized macrocyclic dioxotetraamines has been studied by means of a stopped-flow spectrophotometer. The acid dissociation rate follows the law Vd = CcomkK1K2H 2/(1+K1H+K1K2H 2). From the experimental facts we have obtained, the dissociation kinetics are interpreted by a mechanism involving the negatively charged carbonyl oxygen of the complex being rapidly protonated in a pre-equilibrium step, the rate-determining step being intramolecular hydrogen (enolic tautomer) migration (to imine nitrogen). The dissociation rate reached a plateau in the strongly acidic solution. By means of temperature coefficient method, ΔH φ, ΔS φ of the pre-equilibrium step and ΔH, ΔS of the rate-determining step were obtained. The results of 13-membered macrocyclic dioxotetraamines have been discussed. The influence of the substituents to the acid dissociation rates has also been discussed. The Bronsted type linear free energy relationships do also exist in these C-functionalized dioxotetraamine copper(II) complexes.  相似文献   

11.
The values of density (ρ), viscosity (η) and speed of sound (u) have been measured for binary liquid mixtures of γ-butyrolactone (GBL), δ-valerolactone (DVL), and ε-caprolactone (ECL) with N-methylacetamide (NMA) over the whole composition range at T = (303.15 to 318.15) K and atmospheric pressure. From these data, excess molar volume (VE), deviation in viscosity (Δη), and deviation in isentropic compressibility (Δκs), are calculated. The results are fitted to a Redlich–Kister type polynomial equation to derive binary coefficients and standard deviations.  相似文献   

12.
The structure of solid high-conductance potassium electrolytes K1 − x Al1 − x TixO2 (x = 0.1; 0.2) at 25 and 575°C is studied by a powder neutron diffraction analysis with the application of full-profile Rietveld analysis. Inserting titanium ions removes in potassium aluminate the phase transition at 540°C and the conductance anisotropy typical for its low-temperature form. Both structures are identical (fcc lattice, space group Fd3m). Experiment and calculation coincide best under the assumption that the potassium sublattice is disordered. The conductance increase upon inserting ions Ti4+ is due, apart from stabilization of the fcc structure, to formation of additional potassium vacancies and larger channels for the migration of potassium cations (ions Ti4+ are larger than ions Al3+).__________Translated from Elektrokhimiya, Vol. 41, No. 7, 2005, pp. 878–883.Original Russian Text Copyright © 2005 by Burmakin, Voronin, Akhtyamova, Berger, Shekhtman.  相似文献   

13.
Riaz Qadeer 《Adsorption》2005,11(1):51-55
The kinetics of the adsorption of erbium ions on activated charcoal from aqueous solutions has been studied in the temperature range of 10 to 40∘C. It was observed that the diffusion of erbium ions in to the pores of activated charcoal controls the kinetics of the adsorption process, and the values of intra-particle diffusion rate constant, kd (g/g ⋅ min1/2) were evaluated as 0.7 × 10−3 to 1.6 × 10−3 in the temperature range studied. Various thermodynamic parameters Δ H, Δ G and Δ S were also computed from values of the equilibrium constant KC. The results showed that the adsorption of erbium ions on activated charcoal is an endothermic process.  相似文献   

14.
Reversing-pulse electric birefringence (RPEB) of a nearly monodisperse iron(III) hydroxide oxide sample in the β-form (β-FeOOH) was measured at 25°C and at a wavelength of 633 nm in aqueous media in the presence of NaCl. The concentrations of β-FeOOH and added NaCl varied between 0.00111 and 0.0555 g/L and 0.03 and 2.0 mM, respectively. Except for the suspensions with high salt concentrations, each RPEB signal showed a dip or minimum in the reverse process upon electric field reversal, together with a smooth rise in the buildup and a fall in the decay process. The observed signals were analyzed with a new RPEB theory, which takes into account not only the permanent electric dipole moment (μ) but also the root-mean-square ionic dipole moment (m21/2) due to the ion fluctuation in ion atmosphere, in addition to the field-induced electronic (covalent) dipole moment Δα′ E. The results showed that the slowly fluctuating moment of m21/2 is by far the most predominant one for the field orientation of the β-FeOOH particle, though the permanent dipole moment μ may not be completely excluded. The rotational relaxation time of the whole particle was evaluated from the decay signal, while the relaxation time for fluctuating ions was estimated from RPEB signal fitting. The sign of the steady-state birefringence for β-FeOOH suspensions was positive without exception under the present conditions. The birefringence signals in the steady state (δ/d) were proportional to the second power of the applied field strength (E) in the low field region; thus, the Kerr law was verified to hold for β-FeOOH suspensions. The specific Kerr constant was evaluated for each suspension by extrapolating the values of δ/d to zero field (E→0).  相似文献   

15.
Synthesis and stability under electron irradiation of a hollandite structure-type Ba1.16Al2.32Ti5.68O16 ceramic envisaged for radioactive cesium immobilization. Hollandite structure-type BaxCsy(M,Ti)8O16 (x + y < 2, M trivalent cation) ceramics are currently envisaged as a specific waste form for radioactive cesium immobilization. In order to simulate the effect of cesium β decay on this kind of matrix, the structural modifications and the paramagnetic point defects induced by external electron irradiations near room temperature in a simplified Ba1.16Al2.32Ti5.68O16 hollandite composition were studied mainly by EPR and NMR. Modifications of Al3+ and Ti4+ ions' environment were observed and are due to both the formation of oxygen vacancies and to barium ions displacement. Electron (Ti3+) and hole (O2) centres were observed. The stability of these centres was good at room temperature but thermal treatments performed between 50 and 850 °C generated new paramagnetic defects originating from previous defects. These new defects correspond to titanyl-type Ti3+ ions located on grain surface and to oxygen aggregates in their bulk.  相似文献   

16.
In order to better understand the adsorption mechanism of chitosan-coated magnetic nanoparticles modified with α-ketoglutaric acid (α-KA-CCMNPs), the removal of Cu2+ by α-KA-CCMNPs from aqueous solution was investigated in a batch system at 18, 35 and 50 °C. Different experimental approaches were applied to show mechanistic aspects, such as adsorption isotherms, kinetics and thermodynamics studies. Adsorption equilibrium studies showed that Cu2+ adsorption followed Langmuir model. The kinetics of the interactions was best described by pseudo-second-order mechanism. The thermodynamic parameters (ΔG°, ΔH° and ΔS°) analysis predicted that the adsorption process was strongly dependent on temperature of medium, and spontaneous and endothermic process. The XPS combined with FT-IR spectra revealed that N atom of –NH– group and O atom of carboxyl group in α-KA-CCMNPs coordinated with Cu2+. Experimental results from this study provide data that would be required if this heavy metal adsorption system was to be “scaled up” for industrial application.  相似文献   

17.
The predominance fields of ionic fluorine and metal species in F–M n+–H2O systems (M n+ is Al3+, Be2+, Fe3+, Zr4+, or U4+) were calculated and presented graphically as functions of pH. The corresponding diagrams were plotted, generalized, and analyzed, and the conditions were predicted under which the equilibrium in the system in the presence of any metal complexant shifts toward free fluoride ions. The theoretical prediction was confirmed experimentally and a procedure was developed for the potentiometric determination of fluorine in raw minerals without limitations in composition.  相似文献   

18.
Summary The ion exchange selectivity of trivalent metal ions has been determined on titanium antimonate cation exchanger prepared by coprecipitation of antimony to titanium at different mole ratios. The selectivity sequence Al3+<Cr3+<Ga3+<In3+<Fe3+ was found for trivalent metal ions at an initial concentration of 10–4 mol dm–3 in nitric acid media. A high separation factor Ga/Al = KdGa/KdAl, 4.8×103, was observed for the Ga3+–Al3+ pair on titanium antimonate with an antimony to titanium ratio of 0.34. The effective separation of Ga3+ and In3+ from Al3+ was achieved using a 3 cm×0.5 cm i.d. column containing titanium antimonate with an antimony to titanium ratio of 0.34.  相似文献   

19.
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   

20.
The comparative interfacial oxidation kinetics of the approximate structural isomers trans-(O)2ReV(py)+4 and cis-(O)2ReV(bpy)(py)+2 (py, pyridine; bpy, 2,2′-bipyridine) have been assessed in aqueous solution via conventional cyclic voltammetry at a highly ordered pyrolytic graphite (HOPG) electrode. HOPG was employed because of its known propensity to diminish interfacial electron transfer (ET) rates (by ca. three to four orders of magnitude) and because of a probable lack of importance of kinetic work terms (diffuse double-layer corrections). Measured rates for the trans complex exceed those for the cis by about a factor of 3. Expressed as an effective activation Gibbs energy difference ΔG*, this corresponds to a cis-trans difference of ca. 3 kJ mol−1. The actual vibrational barriers to ET have determined from a combination of published X-ray structural results (trans complex) and new resonance Raman results (cis complex). The values are 0.6 kJ mol −1 for the trans oxidation and 4.4 kJ mol−1 for the cis oxidation (i.e. close to the barrier difference inferred from rate measurements). Further analysis shows that most of the barrier difference is associated with displacement of a (predominantly) Re-N(bpy) stretching mode found only in the cis system. Differences in metal-oxo displacements (cis > trans) are also implicated.  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