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.
The effect of binding by α-alumina is examined on the luminescence spectrum of the uranyl ion. At relatively low concentrations, luminescence studies are complicated by the overlap of the uranyl luminescence and the relatively intense scattering bands of the α-alumina. The presence of uranyl ion is shown to lead to a considerable reduction in the intensity of this scattering, and comparison of emission spectra of this sample with those of alumina, uranyl nitrate and an alumina/uranyl nitrate suspension, followed by computer analysis of the common bands allows an idea of the probable uranyl emission spectrum under these conditions. At high uranyl concentrations less interference from the scattering bands is observed. X-ray diffraction of this sample shows uranyl binding without changing the alumina structure, but with modifications in the lattice dimensions. The room temperature emission spectrum of this sample is very similar to that of uranyl nitrate hexahydrate. In addition, certain other bands are observed which are probably due to Raman scattering. Isopropanol is shown to quench this emission, indicating that solvent access is possible to the uranyl binding site. Luminescence lifetime studies indicate a non-exponential decay, and suggest the possibility of uranyl ion being bound to more than one site. Support for this comes from emission spectra obtained at 77K.  相似文献   

2.
The photoluminescence spectra of the title compounds have been studied under conditions of high resolution at cryogenic temperatures. The luminescence obtained from the 2-methyl imidazolium salt was found to be well behaved, in that all emission originated from intrinsic uranyl centers and all luminescence decay curves consisted of single exponentials. In contrast, the emission associated with the imidazolium salt was found to originate both from intrinsic uranyl groups and from luminescence traps. The analysis of the luminescence decay curves obtained at cryogenic temperatures revealed that the lifetimes characteristic of the trap emission were significantly shorter than those associated with the intrinsic luminescence. The magnetic dipole allowed origins were found to be split by the rhombic crystal field in each system, but this splitting was observed to be 11 cm−1 or less. All other features present in either spectrum were assigned as vibronic in nature and they represent coupling of the various O-U-O or U-Cl vibrational modes with the pure electronic origins.  相似文献   

3.
辜家芳  陈文凯 《物理化学学报》2014,30(10):1810-1820
采用周期性密度泛函理论研究羟基化α-石英(101)面的铀酰离子吸附行为.通过对铀酰离子的水合作用考虑水溶剂对结构的短程溶剂化效应,并通过类导体屏蔽模型(COSMO)考虑水溶剂对结构的远程溶剂化效应.吸附能计算结果和电子结构数据均表明水合铀酰离子吸附构型比氢氧化铀酰吸附构型稳定,并且在液相中两种类型的稳定吸附位均为dia-Os1Os2位.两种形式在电子结构上有很大的差异,主要是由于铀与表面作用后成键强弱程度不同,使5f轨道宽化和略微红移存在差异.在铀酰离子吸附的基础上利用卤素离子改变铀酰离子配位环境可调整体系的带隙.  相似文献   

4.
铀酰离子发光双指数衰减的研究   总被引:1,自引:0,他引:1  
本文系统地研究了铀酰离子发光双指数衰减时的时间分辨发光光谱, 测定和研究了相应条件下铀(Ⅵ)的激光拉曼光谱. 通过计算机对所得光谱进行定量的拟合分解表明, 铀(Ⅵ)的发光双指数衰减是两种发光体UO_2~(2+)和它的水解产物(UO_2)_2(OH)_2~(2+)发光叠加的结果。  相似文献   

5.
辜家芳  陈文凯 《物理化学学报》2001,30(10):1810-1820
采用周期性密度泛函理论研究羟基化α-石英(101)面的铀酰离子吸附行为. 通过对铀酰离子的水合作用考虑水溶剂对结构的短程溶剂化效应,并通过类导体屏蔽模型(COSMO)考虑水溶剂对结构的远程溶剂化效应. 吸附能计算结果和电子结构数据均表明水合铀酰离子吸附构型比氢氧化铀酰吸附构型稳定,并且在液相中两种类型的稳定吸附位均为dia-Os1Os2位. 两种形式在电子结构上有很大的差异,主要是由于铀与表面作用后成键强弱程度不同,使5f 轨道宽化和略微红移存在差异. 在铀酰离子吸附的基础上利用卤素离子改变铀酰离子配位环境可调整体系的带隙.  相似文献   

6.
Luminescence spectra and luminescence decay kinetics of uranyl sulphate water and uranyl nitrate acetone solutions of different concentrations have been studied. Similar experiments have been done with uranyl sulphate powder under vacuum. It has been experimentally shown that the hydrolysis of uranyl sulphate in water takes place, and under low salt concentrations (0.1-4.0 times 10-4 M) a luminescence of a basic form of the photoexcited ion with a tentative structure of UO2OH+* has been observed. The luminescence of the acidic form UO+* has been observed under higher salt concentrations (1–4 times 10-2 M) in water and under any salt concentration in acetone. The acidic form has the characteristic emission spectrum possessing vibrational structure. The luminescence concentrational quenching of both photoexcited uranyl forms and exciplex emission have not been observed. The effect of a number of organic quenchers and molecular oxygen on uranyl luminescence has been studied. There is no luminescence quenching by O2 up to 2 times 106 Pa (20 atm) pressure. The low effectiveness of energy transfer from the photoexcited uranyl forms has been explained in terms of strong steric screening of 5f-uranium (VI) orbital by oxygen atoms and by external filled up uranium electronic shells.  相似文献   

7.
Uranyl ion is photochemically reduced to uranium(IV) in the presence of diphenylmethane and triphenylmethane. Quantum yields for uranium(IV) formation are accelerated with time suggesting the free radical formation, which triggers off a secondary reaction. Lower quantum yield and higherK sv value for photochemical reduction of uranyl ion with triphenylmethane relative to the respective values observed with diphenylmethane reveals the competition between photophysical and photochemical deactivation of excited uranyl ion due to the presence of three and two phenyl groups in respective aromatic hydrocarbons for photophysical deactivation.  相似文献   

8.
The coordination ability of the hexaphosphinoylated p-tert-butylcalix[6]arene B6bL6 toward actinides is established, as well as its good separation ability of the actinide ions UO2 2+ and Th(IV) over trivalent rare earths such as La(III), Eu(III), and Y(III). Spectrophotometric titration of uranyl with B6bL6 in CH 3CN yields log beta 11 = 7.1 and log beta 12 = 12.5 for the 1:1 and 1:2 (UO2 2+/B 6bL6) species, respectively. Actinide complexes with 1:1 and 1:2 (M/L) stoichiometries are isolated and characterized by elemental analysis, IR, and UV-vis. Compounds 1 and 3 fulfill their CN = 8 just with B 6bL (6), while compounds 2 and 4 require coordinated nitrates and/or water molecules. The luminescence spectra of the uranyl complexes and the parameters such as FWMH, vibronic spacing (upsilon sp), and the U-O bond length, as well as the luminescence lifetimes, permit the understanding of the coordination chemistry of these actinide calixarene complexes. Energy transfer from the B6bL6 ligand to the uranyl ion is demonstrated to be relevant in compound 1 with Q abs = 2.0%. The uranyl complex emission reveals a biexponential decay with tau s from 210 to 220 micros and tau L from 490 to 650 micros for compounds 1 and 3, respectively. The liquid-liquid extraction results demonstrate the good extraction capability of B 6bL (6) toward actinides but not for rare earths at room temperature. The extracted species keeps the 1(cation)/1(calixarene) ratio for the UO2 2+, Th 4+, and Eu 3+ ions. A good capacity of B6bL 6 toward Th4+ ions using aqueous phase 2 containing even up to 0.3 M thorium nitrate and an organic phase of 2.47 x 10 (-4) M B6bL6 in chloroform is found. The spectroscopic properties of the isolated uranyl complexes and the extraction studies reveal a uranophilic nature of B6bL6. The molecular modeling results are in good agreement with the experimental findings.  相似文献   

9.
Thiosemicarbazide, phosphoric acid and amidoxime derivatives of chitosan were synthesized and their ability for metal ion adsorptions was discussed. Thiosemicarbazide derivative, synthesized by treating chlorodeoxychitosan with ammonium thiocyanate followed by treatment with hydrazine, was considered to have cross-linked network structure. Phosphoric acid derivative containing both N-phosphonic acid and phosphoric acid groups was synthesized by cyanoethylation of chitosan using acrylonitrile, followed by treatment with hydroxylamine. These derivatives were found to adsorb effectively infinitesimal concentration (ppb order) of uranyl ion in seawater. Stability constants of some metal ion chitosan chelates were determined. To improve the selectivity in the adsorption of metal ions, a novel method utilizing metal ion as a template was adopted, and the results are discussed.  相似文献   

10.
The photoluminescence (PL) of barite is a noncharacteristic property and cannot be used for the investigation of its structure. After thermal treatment of barite at 600°C several luminescent centers were observed, providing information about different impurities. UO 2 2+ was determined from the vibrational structure and the long decay time of the luminescence band. Two different types of uranyl were detected, thin films of uranyl mineral (most probably, reserfordin) and a solid solution of uranyl ion in barite crystal. Characteristic green luminescence of UO 2 2+ may be used as indicative feature for the prospecting of uranium deposits and for the sorting of barite ores with the aim of cleaning from harmful U impurities. Eu2+ was determined from the spectral position, the half-width and the characteristic decay time of the luminescence band. Mn2+ and Ag+ were determined by comparing luminescence bands spectral parameters to those of synthesized BaSO4−Mn and BaSO4−Ag. Fe3+ or Mn4+ were determined from the spectral-kinetic parameters of the luminescence bands. Dedicated to Professor Lisa Heller-Kallai on the occasion of her 65th birthday  相似文献   

11.
The photochemistry and photophysics of aqueous solutions of uranyl nitrate have been investigated by nanosecond laser photolysis with excitation at 266 and 355 nm and by time-resolved fluorescence spectroscopy. The quantum yield has been determined for (UO22+)* formation under excitation with λ = 266 and 355 nm light (φ = 0.35). The quantum yield of uranyl luminescence under the same conditions is 1 × 10–2 and 1.2 × 10–3, respectively, while the quantum yield of luminescence in the solid state is unity, irrespective of the excitation wavelength. The decay of (UO22+)* in the presence of ethanol is biexponential. The rate constants of this process at pH 3.4 are k1 = (2.7 ± 0.2) × 107 L mol–1 s–1 and k2 = (5.4 ± 0.2) × 106 L mol–1 s–1. This biexponential behavior is explained by the existence of different complex uranyl ion species in the solution. The addition of colloidal TiO2 to the solution exerts no effect on the quantum yield of (UO22+)* formation or on the rate of the reaction between (UO22+)* and ethanol. The results of this study have been compared with data available from the literature.  相似文献   

12.
We used molecular dynamics simulations to study the adsorption of aqueous uranyl species (UO(2)(2+)) onto clay mineral surfaces in the presence of sodium counterions and carbonato ligands. The large system size (10,000 atoms) and long simulation times (10 ns) allowed us to investigate the thermodynamics of ion adsorption, and the atomistic detail provided clues for the observed adsorption behavior. The model system consisted of the basal surface of a low-charge Na-montmorillonite clay in contact with aqueous uranyl carbonate solutions with concentrations of 0.027 M, 0.081 M, and 0.162 M. Periodic boundary conditions were used in the simulations to better represent an aqueous solution interacting with an external clay surface. Uranyl adsorption tendency was found to decrease as the aqueous uranyl carbonate concentration was increased, while sodium adsorption remained constant. The observed behavior is explained by physical and chemical effects. As the ionic strength of the aqueous solution was increased, electrostatic factors prevented further uranyl adsorption once the surface charge had been neutralized. Additionally, the formation of aqueous uranyl carbonate complexes, including uranyl carbonato oligomers, contributed to the decreased uranyl adsorption tendency.  相似文献   

13.
To find new adsorbents for uranyl ions, the density functional theory (DFT) was adopted to design a series of new ligands containing an anthracene and two five‐membered heterocycles with nitrogen family nonmetal elements (N, P, As) or oxygen family nonmetal elements (O, S, Se, Te), for example, ligands N,N′‐bis(2‐five‐membered heterocyclidene)‐1,8‐anthradiamines (BFHADAs). Then the uranyl ions were coordinated with BFHADAs to generate five new coordination complexes (Uranyl‐BFHADAs) with heteroatoms N, S, As, Se and Te, respectively. The five‐membered heterocyclic rings of Uranyl‐BFHADA with oxygen atoms were broken under the structural optimization and Uranyl‐BFHADA with heterocyclic atoms P was not obtained. Several structures and property parameters of the ligands BFHADAs (containing heteroatoms N, S, As, Se and Te) and their uranyl complexes Uranyl‐BFHADAs were theoretically investigated and analyzed. The results showed that uranyl ions could form stable coordination complexes with these five BFHADAs. The formed bonds between uranyl ions and the heteroatoms in BFHADAs were coordination bonds rather than other types of bonds. These results could provide insightful information and theoretical guidance for the coordination of uranyl with the atoms N, S, Se, As and Te in other ligands.  相似文献   

14.
Uranyl ion is photochemically reduced to uranium(IV) in the presence of triethylamine and triethylamine is oxidized to secondary amine and acetaldehyde. On the basis of product analysis, temperature independent quantum yields for uranium(IV) formation and abnormal Stern-Volmer plots rule out the simple collisional photochemical annihilation of excited uranyl ion with triethylamine. Static annihilation has a significant contribution in addition to dynamic annihilation.  相似文献   

15.
The effect of gel composition, absorbed dose and pH of the solution on the uranyl ion uptake capacity of N-isopropylacrylamide/maleic acid copolymeric hydrogels containing 0–3 mol% of maleic acid at 48 kGy have been investigated. Uranyl uptake capacity of hydrogels are found to increase from 18.5 to 94.8 mg [UO22+]/g dry gel as the mole % of maleic acid content in the gel structure increased from 0 to 3. The percent swelling, equilibrium swelling and diffusion coefficient values have been evaluated for poly(N-isopropylacrylamide/maleic acid) hydrogels at 500 ppm of uranyl nitrate solution.  相似文献   

16.
Uranyl chelate of 7-iodo-8-hydroxyquinoline-5-sulphonic acid (ferron) has been studied spectrophotometrically in aqueous solution at 25° and at an ionic strength of 0.1 M. The formation of this chelate was pH dependent, and the optimum pH range was between 5.4 to 5.9, Its mole ratio of ligand to uranyl ion was found to be 2 to 1 stoichiometry and the stability constant, log K, was determined as 13.32±0.08. By using the wave-length of 365 mu, determination of trace amount of uranyl ion with the sensitivity of 0.54 γ/cm2 was possible.  相似文献   

17.
The enthalpy which appears during the extraction of a metal ion with an organic solvent can be determined by calorimetric measurements. This measurements were based on the distribution of bismuth, mercury, zinc and cadmium from a solution containing lithium iodide and the extraction of iron, zinc, and uranium from a solution containing potassium thiocyanate. Uranyl nitrate was also investigated. The organic phase consisted of various esters of phosphoric and phosphonic acid and tri-n-octylphosphine oxide in i-octane. The calculated enthalpies are in accordance with the extraction behaviour of the metal ions observed by analytical methods.  相似文献   

18.
Luminescence from aqueous uranyl ions is examined by means of a fluorescence spectrometer with a pulsed xenon light source. Background fluorescence is reduced by using time-based discrimination of the uranyl emission, but interference can still occur from quenchers such as iron(III). Such interferences are reduced by extraction of the uranyl ion into hexane containing trin-n-butyl phosphate, with back-extraction into dilute phosphoric acid before measurement. A detection limit of 5 ng ml?1 is found with a linear calibration range of 0–10 μg ml?1.  相似文献   

19.
Thuéry P 《Inorganic chemistry》2007,46(6):2307-2315
Uranyl nitrate reacts with citric (H4cit) or d-(-)-citramalic (H3citml) acids under mild hydrothermal conditions and in the presence of diamines to give different complexes which are all characterized by the presence of 2:2 uranyl/polycarboxylate dianionic dimers or of polymeric chains based on the same dimeric motif. Each uranium ion is chelated by the two ligands through the alkoxide and the alpha- or beta-carboxylate groups, the second beta-carboxylic group in citrate being uncoordinated. The uranium coordination sphere is completed by either a water molecule or the beta-carboxylate group of a neighboring unit, thus giving zero- or one-dimensional assemblages, respectively. The evidence for [UO2(Hcit)]2 dimers in the solid state confirms previous results from potentiometric and EXAFS measurements on solutions. Depending on the diamine used (DABCO, 2,2'- and 4,4'-bipyridine, [2.2.2]cryptand) and its ability to form divergent hydrogen bonds or not, different uranyl/polycarboxylate topologies are obtained, thus evidencing template effects, and extended hydrogen bonding gives two- or three-dimensional assemblages. These results, together with those previously obtained with NaOH as a base, add to the knowledge of the uranyl/citrate system, which is much investigated for its environmental relevance.  相似文献   

20.
The trinuclear [UO2L]36- and tetranuclear [UO2L]48- metallamacrocycles, obtained by reaction of uranyl nitrate with the rigidly angular ligand (2R,3R,4S,5S)-tetrahydrofurantetracarboxylic acid (H4L) in a basic medium, are in equilibrium in methanol solution. Depending on the counterion, one or the other can be selectively isolated in crystal form. These rare examples of supramolecules incorporating actinide ions confirm the high potential and unique features of uranyl as a building block. Uranyl complexation through both the tri- and bidentate sites of the ligand is at variance with previous assumptions resulting from molecular modeling in nuclear waste reprocessing studies.  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