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 78 毫秒
1.
A method is presented for the cloud-point extraction and separation of copper and lanthanoid ions. A water-soluble calixarene, p-sulfonatocalix[4]arene (C4AS), is used as the chelating agent and Triton X-100 is chosen as the surfactant. The factors affecting the extraction efficiency, such as pH, the concentrations of Triton X-100 and C4AS, equilibration time and centrifugation time, were evaluated. The results demonstrate that there are different extraction behaviors for Cu(II) and Ln(III). Cu(II) can be separated from Ln(III) using C4AS as the chelating agent under weakly acidic conditions. The method may be used to remove trace copper from the lanthanoids.  相似文献   

2.
The adsorption of lanthanides (except for Pm) on mordenite was investigated under various solution conditions of nitrate ion concentrations ([NO*3]: 0.001-2 mol/dm3) and total lanthanide concentrations (0.0005 mol/dm3). Solutions of lanthanide nitrates were equilibrated with zeolite samples at 296 K. A concave tetrad effect was evident in the change of logK d values within the lanthanide series and an explanation by a comparison of covalence in Ln-O bonds existing in triple bond Al-O(1/3Ln)-Si species found in the zeolite phase and in Ln(H2O)3+x or Ln(NO3) n-3 n complexes formed in the aqueous phase is presented. The decreasing trend in C1 and C3 coefficients, which are the function of E1 and E3 Racah f-interelectron repulsion parameters, is evidence of the magnification of covalence in Ln-O bond in the series triple bond S-iO(1/3Ln)-Al triple bond 相似文献   

3.
Adsorption of atrazine on soils: model study   总被引:1,自引:0,他引:1  
The adsorption of the widely used herbicide atrazine onto three model inorganic soil components (silica gel, gamma-alumina, and calcite (CaCO(3)) was investigated in a series of batch experiments in which the aqueous phase equilibrated with the solid, under different solution conditions. Atrazine did not show discernible adsorption on gamma-alumina (theta=25 degrees C, 3.8相似文献   

4.
The adsorption of bovine serum albumin (BSA) at the air-water interface has been studied by specular neutron reflection. The variation of the adsorbed amount and the total thickness of the BSA layer with respect to bulk BSA concentration was determined at pH 5, close to its isoelectric point (IP). While the surface excess showed a steady increase with bulk concentration the thickness of the protein layer was found to be close to the short axial length of 40 ? of the globular solution structure of BSA at concentrations below 0.1 g dm-3, suggesting that BSA molecules adsorb with their long axes parallel to the surface of water. At 1 g dm-3 the adsorbed layer can be modeled as an upper layer of 40 ? with a volume fraction of 0.4 and a sublayer of 30 ? underneath the top main layer with a volume fraction of 0.12. The results suggest that, although there is some structural deformation accompanying adsorption, there is no denaturation. The extent of immersion of the BSA in water was determined by performing the measurements in D2O and in a mixture of H2O and D2O whose contrast matches that of BSA. The signal is then only from the part of the layer out of water. At pH 5 this layer was about 10 +/- 5 ? at a bulk concentration of 5 x 10(-4) g dm-3 and decreased to 5 +/- 3 ? at 1 g dm-3. The fraction of the BSA layer immersed in water therefore varies from about 70 to over 90%. The effect of pH on the adsorption was examined at two BSA concentrations. While pH had little effect on the adsorption at a low BSA concentration of 5 x 10(-3) g dm-3, both surface excess and layer thickness showed pronounced peaks at pH 5 at the higher concentration of 1 g dm-3. The increased adsorption at pH 5 is attributed to the reduced lateral electrostatic repulsion around the IP. This adsorption pattern became less pronounced when the total ionic strength was increased from 0.02 to 1 M, indicating that the electrolyte screens the electrostatic repulsions within the adsorbed layer. Copyright 1999 Academic Press.  相似文献   

5.
A chelating agent-loaded resin consisting of 8-quinolinol-5-sulfonic acid and an anion-exchange resin (HOx-resin) was prepared in order to concentrate trace chalcophile elements in natural water samples selectively before neutron activation analysis. The exchange capacity of the Diaion SA No. 100 for the reagent (1.8 meq . g-1 resin) corresponds approximately to that for chloride ion (1.83 meq . g-1 resin), indicating that 8-quinolinol-5-sulfonic acid is adsorbed quantitatively on the exchange site of the resin through the sulfonate anion in the reagent. The basic conditions for the adsorption of the metal ions on the resin were investigated by employing the column method. The nitrate concentration and the pH of the sample solution affect the adsorption behavior of metal ions. Several solutions containing metal ions with varying pH or varying nitrate concentration were applied to the resin column (35 mm x 7 mm phi) with a flow rate of 2.0 cm3 . min-1. As a result, the optimum conditions for the quantitative adsorption of copper(II), zinc(II), cadmium(II), cobalt(II), nickel(II) and manganese(II) were as follows: NO3- less than 0.01 mol . dm-3 pH greater than 4.6. Furthermore, the feasibility of the above conditions as well as quantitative adsorption of the chalcophile elements was confirmed through the neutron activation analysis of the synthesized metal solutions.  相似文献   

6.
The rates of dissolution of calcitic Carrara marble have been reported to be significantly reduced in alkaline pH (pH 8.25) at 25 degrees C in the presence of (1-hydroxyethylidene)-1,1 diphosphonic acid (HEDP). The adsorption takes place at the calcite/water interface at the double layer through the interaction of charged surface species with the charged solution species of the adsorbate. The present work focused on obtaining a better understanding of the interaction of the calcite surface with HEDP. Calculations were performed according to the triple layer model, assuming the formation of surface complexes between the charged surface species of calcite and the species of HEDP dominant at pH 8.25. According to the model, the adsorbed species are located at the inner Helmholtz plane of the electrical double layer. Strong lateral interactions between the adsorbed species were suggested and were corroborated from the calculation of the respective energy, which was equal to 69 kJ mol(-1). The adsorption isotherm was consistent with the proposed model at low surface coverage values, while discrepancies between the values experimentally measured and the predicted were found at higher adsorbate concentrations. The deviations from the predicted values were attributed to the fact that HEDP adsorption on calcite resulted in the formation of multiple layers. The model explained adequately the changes in the zeta-potential values of calcite in the presence of HEDP in the solution which resulted in charge reversal upon adsorption.  相似文献   

7.
锗(IV)-茜素红-VOSO4-EDTA体系极谱吸附催化波研究   总被引:1,自引:0,他引:1  
李南强  崔毅  何红兵 《化学学报》1987,45(4):399-402
报导了在酸性缓冲液中锗(IV)-茜素红-VOSO4-EDTA体系的配合吸附平行催化波的研究.锗的最低检出量为1.0x10^-^9mol.dm^-^3.研究了催化波的机理.  相似文献   

8.
The solvent extraction of fourteen lanthanoid ions with thenoyltrifluoroacetone (HTTA) in combination with tetraethyl 4-tert-butylcalix[4]arene-tetraacetic acid tetraethyl ester (S) from a perchlorate medium at constant ionic strength was investigated. The extracted species were identified as the Ln(TTA)3·S complexes by slope analysis. Equilibrium constants, parameters for extraction, and the synergistic and separation factors between two adjacent Ln(III) ions were determined.
Online abstract
The solvent extraction of 14 lanthanoids (Ln) with thenoyltrifluoroacetone (HTTA) in combination with 4-tert-butylcalix[4]arene-tetraacetic acid tetraethyl ester (S) is investigated. The composition of the extracted species has been found as Ln(TTA)3·S. The values of the equilibrium constants, the synergistic and separation factors between two adjacent Ln(III) ions have been calculated.  相似文献   

9.
Samples of polycrystalline calcite were impregnated with solutions of malonic acid of three concentrations (5 x 10(-2), 5 x 10(-3) , and 5 x 10(-4) M) and different pH values (6.00, 7.00, and 8.00). The impregnation was carried out at room temperature to evaluate the adsorption of malonate ion in the calcite surface to optimize the conditions for possible application on limestone and marble in cultural heritage materials. The affinity of the malonate ion was determined through the potentiometric measurement of the surface charge and the corresponding adsorbed amounts by titration, Raman spectroscopy, and small-angle X-ray scattering (SAXS). The results indicate effective adsorption of the malonate ion on the surface at a pH value close to the point of zero charge (pHpzc approximately 8.20) and changes in some surface morphological properties such as the pore shape and the pore size distribution. The presence of a malonate adsorptive layer on calcite generates an interface interaction potential that may influence the reaction and transport mechanisms within the medium.  相似文献   

10.
A study of competitive adsorption of Ca(2+) and Zn(II) ions at the monodispersed SiO(2)/electrolyte solution interface is presented. Influence of ionic strength, pH, and presence of other ions on adsorption of Ca(2+) and Zn(II) in the mentioned system are investigated. zeta potential, surface charge density, adsorption density, pH(50%), and DeltapH(10-90%) parameters for different concentrations of carrying electrolyte and adsorbed ions are also presented. A high concentration of zinc ions shifts the adsorption edge of Ca(2+) ions adsorbed from solutions with a low initial concentration at the SiO(2)/NaClO(4) solution interface to the higher pH values. This effect disappears with a concentration increase of calcium ions. The presence of Ca(2+) ions in the system slightly affects the adsorption of zinc ions on SiO(2), shifting the adsorption edge toward lower pH values and thereby increasing the adsorption slope.  相似文献   

11.
This is the first systematic study exploring the potential of high-energy EXAFS as a structural tool for lanthanoids and third-row transition elements. The K-edge X-ray absorption spectra of the hydrated lanthanoid(III) ions both in aqueous solution and in solid trifluoromethanesulfonate salts have been studied. The K-edges of lanthanoids cover the energy range from 38 (La) to 65 keV (Lu), while the corresponding energy range for the L(3)-edges is 5.5 (La) to 9.2 keV (Lu). We show that the large widths of the core-hole states do not appreciably reduce the potential structural information in the high-energy K-edge EXAFS data. Moreover, for lanthanoid compounds, more accurate structural parameters are obtained from analysis of K-edge than from L(3)-edge EXAFS data. The main reasons are the much wider k range available and the absence of double-electron transitions, especially for the lighter lanthanoids. A comparative K- and L(3)-edge EXAFS data analysis of nonahydrated crystalline neodymium(III) trifluoromethanesulfonate demonstrates the clear advantages of K-edge analysis over conventionally performed studies at the L(3)-absorption edge for structural investigations of lanthanoid and third-row transition metal compounds. The coordination chemistry of the hydrated lanthanoid(III) ions in aqueous solution and solid trifluoromethanesulfonate salts, based on the results of both the K- and L(3)-edge EXAFS data, is thoroughly discussed in the next paper in this series (I. Persson, P. D'Angelo, S. De Panfilis, M. Sandstr?m, L. Eriksson, Chem. Eur. J. 2008, 14, DOI: 10.1002/chem.200701281).  相似文献   

12.
The liquid extraction of 14 lanthanoids with a 4-benzoyl-3-phenyl-5-isoxazolone (HPBI) alone in CHCl3 as a diluent from perchlorate medium at constant ionic strength μ = 0.1 is investigated. The synergistic solvent extraction of five selected lanthanoid ions (La3+, Nd3+, Eu3+, Ho3+ and Lu3+) with 4-benzoyl-3-phenyl-5-isoxazolone (HPBI) and 5,11,17,23-tetra-tert-butyl-25,26,27,28-tetrakis-(dimethylphosphinoylmethoxy)calix[4]arene, (S) in CHCl3 has been studied too. The stoichiometry of the extracted species was characterised by a classical log–log plot analysis. It was found that the composition of the extracted species with HPBI are Ln(PBI)3 and in the presence of the phosphorus-containing calix[4]arene the lanthanoid ions have been extracted as [Ln(PBI)3S2]. The values of the equilibrium constants and the separation factors have been calculated. The influence of the synergistic agent on the extraction process has been discussed.  相似文献   

13.
The title subject has been studied by galvanostatic single-pulse, chronopotentiometric and equilibrium measurements on the Zn(Hg)/Zn(II) electrode in x M KI+(1?x) M KCl (x from 0 to 1), 1 M KBr and 1 M MeCl (Me=Li, Na, K and Cs) solutions of pH 3 at 25°C. Quantitative information about the effect of specifically adsorbed halides on the rates of the Zn(II)/Zn(I) and the Zn(I)/Zn(Hg) steps is obtained separately (for the latter step mainly at potentials near ?1.0 V(SCE)), and the latter step seems to be more influenced than the former by the adsorption. An attempt is made to correlate the adsorption effect on the rate of the Zn(II)/Zn(I) step to double-layer parameters according to recent models for such effects. The extra current observed at potentials where the halides are adsorbed, seems to vary with the surface activity of the specifically adsorbed ion. The lack of any observed kinetic effect of Cs+, which is specifically adsorbed at these potentials, is possibly due to the Cs+ specific adsorption enhancing the Cl? specific adsorption and vice versa, so that the decelerating and accelerating effects by these ions may cancel each other.  相似文献   

14.
《Analytical letters》2012,45(1):157-168
Abstract

4-(2-Thiazolylazo)-6-chlororesorcinol (TAR-Cl) reacts sensitively with uranyl(II) and lanthanoids(III), and forms reddish-brown 1:1 and 1:2 complexes. The complexing behaviors were examined spectrophotometrically. The absorption maxima of the complexes are focused near 553 nm and the optimum pH for complexation lies between 6.5–8.8. Beer's law holds up to about 2 × 10?5 mol 1?1, with a molar absorptivity of 3.00 × 104 1 mol?1 cm?1 for uranium and 6 × 104 1 mol?1 cm?1 level for each lanthanoid. The absorptivities are increased with the atomic number, especially in light lanthanoids, that are correlative both to the lanthanoid contraction and the basicity of ortho hydroxyl group in the resorcinol ring, but such effects are not clearly recognized in heavy lanthanoids. Effect of masking agents was also examined, and uranium could be determined selectively in the presence of lanthanoid mixtures by the addition of CyDTA.  相似文献   

15.
Single‐, double‐, and multi‐walled carbon nanotubes (SWCNTs, DWCNTs, and MWCNTs), and two oxidized MWCNTs with different oxygen contents (2.51 wt % and 3.5 wt %) were used to study the effect of the wall number and surface functionalization of CNTs on their adsorption capacity and adsorption–desorption hysteresis for heavy metal ions (NiII, CdII, and PbII). Metal ions adsorbed on CNTs could be desorbed by lowering the solution pH. Adsoprtion of heavy metal ions was not completely reversible when the supernatant was replaced with metal ion‐free electrolyte solution. With increasing wall number and amount of surface functional groups, CNTs had more surface defects and exhibited higher adsorption capacity and higher adsorption–desorption hysteresis index (HI) values. The coverage of heavy metal ions on the surface of CNTs, solution pH, and temperature affect the metal ion adsorption–desorption hysteresis. A possible shift in the adsorption mechanism from mainly irreversible to largely reversible processes may take place, as the amount of metal ions adsorbed on CNTs increases. Heavy metal ions may be irreversibly adsorbed on defect sites.  相似文献   

16.
Phenalen-1-one was obtained in considerable yield by reducing 3-hydroxyphenalen-1-one. Most of known preparation methods are not very practical, either because their yields are very poor or because their processes have many steps. This regioselective 1,2-reduction proceeded by the action of NaBH4 and various cations of rare-earth elements and metals. The yields of phenalen-1-one were examined as a function of typical lanthanoids, molar ratios of lanthanoid ions to 3-acetoxy-phenalen-1-one, and differing methods of protecting the hydroxyl group. Lanthanum chloride (LaCl3) gave the greatest yield (45.3%) of phenalen-1-one at molar ratios higher than a third, probably because La3+ ion is a hard acid and coordinates easily to a hard solvent such as methanol. Further, it has the largest ionic radius among all lanthanoid ions.  相似文献   

17.
Coordination of nitrate anions with lanthanoid cations (Ln(3+)) in water, methanol and octanol-1 has been studied by means of molecular dynamics simulations with explicit polarization. Potential of mean force (PMF) profiles have been calculated for a mono-complex of lanthanoid nitrate (Ln(NO(3))(2+)) in these solvents using umbrella-sampling molecular dynamics. In pure water, no difference in the nitrato coordination to lanthanoids (Nd(3+), Eu(3+) and Dy(3+)) is observed, i.e. the nitrate anion prefers the monodentate coordination, which promotes the salt dissociation. Then, the influence of the nature of the solvating molecules on the nitrato coordination to Eu(3+) has been investigated. PMF profiles point out that both monodenate and bidentate coordinations are stable in neat methanol, while in neat octanol, only the bidentate one is. MD simulations of Eu(NO(3))(3) in water-octanol mixtures with different concentrations of water have been then performed and confirm the importance of the water molecules' presence on the nitrate ion's coordination mode.  相似文献   

18.
A brine sample and two deposit samples of a salt lake in China were analyzed for their contents of lanthanoids (Ln's), thorium and uranium by neutron activation analysis. Five Ln's were determined at sub ppb levels. Th and U contents were at about the same levels as those of Ln's. The lanthanoid abundance patterns (Ln pattern) of the three samples are similar to each other, each having a negative slope in the light Ln region. There seems no substantial difference in distribution between the solution (brine) and solid (deposit) phases among Ln's.  相似文献   

19.
钴(II)-丁二酮肟-亚硝酸盐体系极谱催化波的机理研究   总被引:1,自引:0,他引:1  
倪亚明  李玲  高小霞 《化学学报》1988,46(7):651-656
在氨性底液(PH8)中, 钴(II)-丁二酮肟(DMG)-亚硝酸盐体系产生高灵敏的极谱催化波. 利用吸附伏安法, 测定下限可达1×10^-^1^1mol.dm^-^3Co. 我们用多种电化学方法和紫外可见分光光度法证明, 吸附在汞电极表面的[NH4]2[Co(DMG)2(NO2)2]是有很高电活性的混配化合物, 在复杂的电还原过程中, 不仅Co(II)和DMG被催化还原, 而且NO2^-也被催化还原, 从而产生很大的催化电流, 本文再一次证明, “活性钴"在催化波的形成过程中起着重要的作用.  相似文献   

20.
The temperature effect on the magnitude of adsorption was used to explain the mechanism of adsorption of gadolinium on silica at very low concentrations. Standard enthalpy of adsorption of gadolinium equals 36 kJ mol-1 for a total Gd concentration of 2 x 10(-8) mol dm-3 and 67 kJ mol-1 for 2 x 10(-5) mol dm-3. This result confirms the hypothesis that the Gd adsorption at low initial concentration is governed by formation of strong ternary surface complexes involving anionic impurities. Copyright 1999 Academic Press.  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