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.
Vibrational spectra have been obtained for aqueous solution of uranyl-perchlorates, -fluorides, -chlorides, -acetates and -sulphates over a range of solution composition with added anions. We have prepared [Bun4N][UO2Cl4], [Me4N][UO2Cl4], [Prn4N][[UO2(NO3)3], [Bun4N][UO2(NO3)3], with the expectation that the large cation would give a better approximation to vibrational frequencies of the free anion and would allow measurements in non-coordinating solvents. As the perchlorate is not coordinated to [UO2]2+ in aqueous solution the expected structure is a solvated cation [UO2(OH2)5]2+ with characteristic infrared 962.5, 253 and 160 cm−1 and Raman 874 and 198 cm−1 bands. The formation of weak, solvated [UO2X]+ complexes (X=F, Cl) has been established with frequencies at 908, 827, 254, 380 cm−1 and 956, 871, 254 and 222 cm−1 for [UO2F]+ and [UO2Cl]+, respectively. Bidentate NO3 coordination has been established for solid and dissolved (in CH2Cl2) [R4N][UO2(NO3)3] (R=Prn, Bun). Aqueous solutions of UO2(NO3)2 and Cs[UO2(NO3)3] show no clear evidence that bidentate or monodentate nitrate is present. Both unidentate and bidentate linkage of acetate-uranyl were established for acetate complexes in aqueous solutions. For the uranyl sulphate system, monodentate sulphate coordination is the major mode at low SO4:U ratios, and even at a ratio of 3:1 there is very little free sulphate.  相似文献   

2.
2,2′-Bipyridine-N,N′-dioxide (bypO2 = L) complexes of the composition [UO2(bypO2)2(NO3)2]·2H2O (UO2–L2–NO3), [UO2(bypO2)2H2O](ClO4)2 (UO2–L2–ClO4) and [UO2bypO2(H2O)2SO4] (UO2–L–SO4) have been prepared by the reactions of the respective hydrated uranyl salts with the bypO2 ligand in water. The structures of the complexes were elucidated using elemental and thermal analyses, IR and luminescence spectroscopy as well as luminescence lifetime measurements. The IR spectra show that the bonding between uranium and bypO2, as well as uranium and water or a counter ion (NO3 and SO42−) is formed. The nitrate or sulfate groups coordinate to the central metal ions in a monodentate manner. From TG–DTA curves, the nature of the water molecules present in the complexes and the decomposition temperature of the dehydrated uranyl complexes were determined. The thermal stability of the anhydrous uranyl complexes increases in the series: (UO2–L2–NO3) < (UO2–L2–ClO4) < (UO2–L–SO4). All the compounds show green-yellow intense luminescence. The main fluorescence bands and the emission lifetimes in these complexes were determined. The luminescence spectra of all the prepared complexes differ from each other with respect to their peak maxima positions. The luminescence lifetimes also vary. The structure of the (UO2–L–SO4) complex was determined by X-ray single-crystal analysis.  相似文献   

3.
UO2+–solvent complexes having the general formula [UO2(ROH)]+ (R=H, CH3, C2H5, and n‐C3H7) are formed using electrospray ionization and stored in a Fourier transform ion cyclotron resonance mass spectrometer, where they are isolated by mass‐to‐charge ratio, and then photofragmented using a free‐electron laser scanning through the 10 μm region of the infrared spectrum. Asymmetric O=U=O stretching frequencies (ν3) are measured over a very small range [from ~953 cm?1 for H2O to ~944 cm?1 for n‐propanol (n‐PrOH)] for all four complexes, indicating that the nature of the alkyl group does not greatly affect the metal centre. The ν3 values generally decrease with increasing nucleophilicity of the solvent, except for the methanol (MeOH)‐containing complex, which has a measured ν3 value equal to that of the n‐PrOH‐containing complex. The ν3 frequency values for these U(V) complexes are about 20 cm?1 lower than those measured for isoelectronic U(VI) ion‐pair species containing analogous alkoxides. ν3 values for the U(V) complexes are comparable to those for the anionic [UO2(NO3)3]? complex, and 40–70 cm?1 lower than previously reported values for ligated uranyl(VI) dication complexes. The lower frequency is attributed to weakening of the O?U?O bonds by repulsion related to reduction of the U metal centre, which increases electron density in the antibonding π* orbitals of the uranyl moiety. Computational modelling of the ν3 frequencies using the B3LYP and PBE functionals is in good agreement with the IRMPD measurements, in that the calculated values fall in a very small range and are within a few cm?1 of measurements. The values generated using the LDA functional are slightly higher and substantially overestimate the trends. Subtleties in the trend in ν3 frequencies for the H2O–MeOH–EtOH–n‐PrOH series are not reproduced by the calculations, specifically for the MeOH complex, which has a lower than expected value.  相似文献   

4.
The crystal structures of two uranyl nitrate complexes with 2,2′-6,2″-terpyridine (Trpy), [(UO (Trpy)-[(UO2)2(Trpy)2(OH)2][UO2(NO3)2(OH)2] · 2CH3OH (I) and [(UO2)2(Trpy)2(OH)2]2[UO2(NO3)3(H2O)](NO3)3 · 3H2O (II), were studied. Compound I consists of the centrosymmetric dimeric cations [(UO2)2(Trpy)2(OH)2]2+, the anions [UO2(NO3)2(OH)2]2?, and solvation methanol molecules. Complex II consists of dimeric cations [(UO2)2(Trpy)2(OH)2]2+, the complex anions [UO2(NO3)3(H2O)]?, nitrate anions, and water molecules of crystallization. The uranium atom in the [UO2(NO3)3(H2O)]? anion in II has an unusual coordination polyhedron representing a hexagonal bipyramid in which one oxygen atom in the equatorial plane is replaced by two atoms equidistant from this plane.  相似文献   

5.
Uranyl–sulphate complexes are the predominant U(VI) species present in acid solutions resulting either from underground uranium ore leaching or from the remediation of leaching sites. Thus, the study of U(VI) speciation in these solutions is of practical significance. The spectra of UO2(NO3)2 + Na2SO4 solutions of different Φ S = [SO42−]/[U(VI)] ratio at pH = 2 were recorded for this purpose. As the presence of uranyl-nitrate complexes should be expected under these experimental conditions, the spectra of UO2(NO3)2 + NaNO3 solutions with different Φ N = [NO3]/[U(VI)] ratio at pH = 2 were also measured. The effects of Φ S and Φ N ratios value were most pronounced in wavelength interval 380–500 nm. Therefore, these parts of experimental overall spectra were used for deconvolution into the spectra of individual species by the method proposed. It enabled to calculate stability constants of anticipated species at zero ionic strength. The Specific Ion Interaction Theory (SIT) was used for this purpose. Stability constants of UO2SO4, UO2(SO4)22−, UO2NO3 + and UO2(NO3)2 coincided well with published data, but those for UO2(SO4)34− and UO2(NO3)3 were significantly lower.  相似文献   

6.
Simple and versatile routes to the functionalization of uranyl‐derived UV–oxo groups are presented. The oxo‐lithiated, binuclear uranium(V)–oxo complexes [{(py)3LiOUO}2(L)] and [{(py)3LiOUO}(OUOSiMe3)(L)] were prepared by the direct combination of the uranyl(VI) silylamide “ate” complex [Li(py)2][(OUO)(N”)3] (N”=N(SiMe3)2) with the polypyrrolic macrocycle H4L or the mononuclear uranyl (VI) Pacman complex [UO2(py)(H2L)], respectively. These oxo‐metalated complexes display distinct U? O single and multiple bonding patterns and an axial/equatorial arrangement of oxo ligands. Their ready availability allows the direct functionalization of the uranyl oxo group leading to the binuclear uranium(V) oxo–stannylated complexes [{(R3Sn)OUO}2(L)] (R=nBu, Ph), which represent rare examples of mixed uranium/tin complexes. Also, uranium–oxo‐group exchange occurred in reactions with [TiCl(OiPr)3] to form U‐O? C bonds [{(py)3LiOUO}(OUOiPr)(L)] and [(iPrOUO)2(L)]. Overall, these represent the first family of uranium(V) complexes that are oxo‐functionalised by Group 14 elements.  相似文献   

7.
Acetylpyridine benzoylhydrazone and related ligands react with common dioxouranium(VI) compounds such as uranyl nitrate or [NBu4]2[UO2Cl4] to form air‐stable complexes. Reactions with 2, 6‐diacetylpyridinebis(benzoylhydrazone) (H2L1a) or 2, 6‐diacetylpyridinebis(salicylhydrazone) (H2L1b) give yellow products of the composition [UO2(L1)]. The neutral compounds contain doubly deprotonated ligands and possess a distorted pentagonal‐bipyramidal structure. The hydroxo groups of the salicylhydrazonato ligand do not contribute to the complexation of the metal. The equatorial coordination spheres of the complexes can be extended by the addition of a monodentate ligand such as pyridine or DMSO. The uranium atoms in the resulting deep‐red complexes have hexagonal‐bipyramidal coordination environments with the oxo ligands in axial positions. The sterical strains inside the hexagonal plane can be reduced when two tridentate benzoylhydrazonato ligands are used instead of the pentadentate 2, 6‐diacetylpyridine derivatives. Acetylpyridine benzoylhydrazone (HL2) and bis(2‐pyridyl)ketone benzoylhydrazone (HL3) deprotonate and form neutral, red [UO2(L)2] complexes. The equatorial coordination spheres of these complexes are puckered hexagons. X‐ray diffraction studies on [UO2(L1a)(pyridine)], [UO2(L1b)(DMSO)], [UO2(L2)2] and [UO2(L3)2] show relatively short U—O bonds to the benzoylic oxygen atoms between 2.328(6) and 2.389(8) Å. This suggests a preference of these donor sites of the ligands over their imino and amine functionalities (U—N bond lengths: 2.588(7)—2.701(6) Å ).  相似文献   

8.
The interactions of yttrium chloride, zirconium chloride, and uranium nitrate with sparfloxacin (SPAR) in ethanol, methanol, and acetone were studied. The isolated solid complexes were characterized by elemental analysis, infrared, 1H-NMR and electronic spectra, and thermogravimetric analysis. The results support the formation of [Y(SPAR)2Cl2]Cl ? 12H2O, [ZrO(SPAR)2Cl]Cl ? 15H2O, and [UO2(SPAR)3](NO3)2 ? 5H2O. Infrared spectra of the isolated solid complexes indicate that SPAR is bidentate through the ring carbonyl oxygen and one oxygen of carboxylate. The calculated bond length and force constant, F(U=O), in the uranyl complex are 1.747 Å and 655.29 Nm?1, respectively. The antimicrobial activities of the ligand and metal complexes have been tested against bacteria Staphylococcus aureus (S. aureus), Escherichia coli (E. coli) and Pseudomonas aeruginosa (P. aeruginosa) and fungi Penicillium rotatum (P. rotatum) and Trichoderma sp., showing that the complexes exhibit higher antibacterial activity than SPAR.  相似文献   

9.
On the basis of uranyl complexes reacting with a polypyrrolic ligand (H4L), we explored structures and reaction energies of a series of new binuclear uranium(VI) complexes using relativistic density functional theory. Full geometry optimizations on [(UO2)2(L)], in which two uranyl groups were initially placed into the pacman ligand cavity, led to two minimum‐energy structures. These complexes with cation–cation interactions (CCI) exhibit unusual coordination modes of uranyls: one is a T‐shaped ( T ) skeleton formed by two linear uranyls {Oexo?U2?Oendo→U1(?Oexo)2}, and another is a butterfly‐like ( B ) unit with one linear uranyl coordinating side‐by‐side to a second cis‐uranyl. The CCI in T was confirmed by the calculated longest distance and lowest stretching vibrational frequency of U2?Oendo among the four U?O bonds. Isomer B is more stable than T , for which experimental tetrameric analogues are known. The formation of B and T complexes from the mononuclear [(UO2)(H2L)(thf)] ( M ) was found to be endothermic. The further protonation and dehydration of B and T are thermodynamically favorable. As a possible product, we have found a trianglelike binuclear uranium(VI) complex having a O?U?O?U?O unit.  相似文献   

10.
By using 2,2′-bipyridine N-oxide (bipyO) and 2,2′-bipyridine N,N′-dioxide (bipyO2), three new uranyl complexes [UO2(bipyO)SO4]·H2O (1), [UO2(bipyO)(OH)(NO3)]2·H2O (2) and [UO2(bipyO2)H2O](ClO4)2·(3) were synthesized using uranyl salts including non-coordinating or weakly coordinating power of the ClO4 anion and the strongly coordinating power of NO3 and SO42− anions. All of the compounds were characterized by CHN microanalytical procedures, infrared and luminescence spectroscopy and by single crystal X-ray diffraction. Spectroscopic studies indicate that the bipyO is bound to the uranyl group via the nitrogen and oxygen atoms. Structural analyses revealed that overall bonding pattern is different in each case: 1 is a polymer; in 2 dimeric complex molecules are formed, whereas 3 is composed of monomers. In all of the complexes, the uranium atom is in a seven-coordinate environment.  相似文献   

11.
The extraction of thorium(IV) and uranium(VI) from nitric acid solutions has been studied using mixtures of bis(2,4,4-trimethylpentyl)phosphinic acid (Cyanex272 or HA), and synergistic extractants (S) such as tri-butylphosphate (TBP), tri-octylphosphine oxide (TOPO) or bis(2,4,4-trimethylpentyl)thiophosphinic acid (Cyanex301). The results showed that these metallic ions are extracted into kerosene as Th(OH)2(NO3)A·HA and UO2(NO3)A·HA with Cyanex272 alone. In the presence of neutral organophosphorus ligands TBP and TOPO, they are found to be extracted as Th(OH)2(NO3)A·HA·S and UO2(NO3)A·HA·S. On the other hand, Th(IV), U(VI) are extracted as Th(OH)2(NO3)A·HA·2S and UO2(NO3)A·HA·S in the presence of Cyanex301. The addition of neutral extractants such as TOPO and TBP to the extraction system enhanced the extraction efficiency of both elements while Cyanex301 as an acidic extractant has improved the selectivity between uranium and thorium. The effect of TOPO on the extraction was higher than other extractants. The equilibrium constants of above species have been estimated by non-linear regression method. The extraction amounts were determined and the results were compared with those of TBP. Also, it was found that the binding to the neutral ligands by the thorium–Cyanex272 complexes follows the neutral ligand basicity sequence.  相似文献   

12.
Uranyl(VI) complexes [UO2(L)(solvent)], where L denotes an asymmetric N2O2 Schiff base (salpyr, 3-MeOsalpyr, 4-MeOsalpyr, 5-MeOsalpyr, 5-Clsalpyr or 5-Brsalpyr; salpyr is N,N′-bis(salicyliden)-2,3-diaminopyridine), were synthesized and characterized in solution (UV–vis, 1H NMR, cyclic voltammetry) and in the solid-state (X-ray crystallography, IR, TGA, C H N.). X-ray crystallography of UO2(3-MeOsalpyr) revealed coordination of the uranyl by the tetradentate Schiff base and one disordered solvent, resulting in seven-coordinate uranium. Another disordered solvent was not coordinated. Cyclic voltammetry of [UVIO2(L)(solvent)] in acetonitrile was used to investigate the effect of the substituents of the Schiff base ligands on oxidation and reduction potential. The quasi-reversible redox reaction without any successive reactions was accelerated by groups with lesser electron withdrawing. We also investigated the kinetics and mechanism of the exchange reaction of the coordinated solvent with tributylphosphine using spectrophotometric method. The second-order rate constants at four temperatures and activation parameters showed an associative mechanism for all corresponding complexes with the following trend: UO2(5-Clsalpyr)?>?UO2(5-Brsalpyr)?>?UO2(3-MeOsalpyr)?>?UO2(4-MeOsalpyr)?>?UO2(salpyr)?>?UO2(5-MeOsalpyr). It was concluded that the steric and electronic properties of the complexes were important for the reaction rate.  相似文献   

13.
The nonlinear optical activity (Q) of uranyl carboxylates containing [UO2(L)3] complexes in the crystal structure, where L is the anion of aliphatic or unsaturated monocarboxylic acids, has been characterized for the first time by the second harmonic generation method. It has been shown using molecular Voronoi–Dirichlet polyhedra that specific features of the cationic sublattice of U and R atoms in carboxylate structures depend on noncovalent interactions between outer-sphere R+ cations and [UO2(L)3] complex anions. The existence of a relationship between Q and the magnitude of the vector characterizing the displacement of the uranium atom nucleus from the centroid of its Voronoi–Dirichlet polyhedron in the cation sublattice of U and R atoms is revealed.  相似文献   

14.
Novel five binuclear polymeric dioxouranium(VI) of azosulphadrugs [(azodrug substances) azobenzene sulphonamides] were prepared for the first time. The infrared spectra of the samples were recorded and their fundamental vibration wave number was obtained. The resulting polymeric uranyl complexes were characterized on the basis of their elemental analyses, conductance and spectral (IR, NMR, and electronic spectra) data. The ligation modes of the azosulphadrugs ligands towards uranyl(II) ions were critically assigned and addressed properly on the basis of their IR and their uranyl(II) complexes. The theoretical aspects are described in terms of the well-known theory of 5d–4f transitions. The coordination geometries and electronic structures are determined from a framework for the modeling of novel polymer complexes. The values of ν3 of the prepared complexes containing UO22+ were successfully used to calculate the force constant, FUO (1n 10?8 N/Å) and the bond length RUO of the U–O bond. Wilson's, matrix method, Badger's formula, and Jones and El-Sonbati equations were used to calculate the U–O bond distances from the values of the stretching and interaction force constants. The most probable correlations between U–O force constant to U–O bond distance were satisfactorily discussed in terms of “Badger's rule”, “Jones” and “El-Sonbati equations”.  相似文献   

15.
A method for selective extraction of uranium from carbonate solutions containing molybdate is reported. A liquid chromatography column, packed with N-β-aminoethyl-γ-ammopropyltrimethoxysilane immobilized on a glass substrate, was utilized in a continuous How system. The selective retention of the uranyl carbonate species [UO2(CO3)2? 2H2O]2- and UO2(CO3)4- on protonated immobilized diamine is the basis for this separation Recoveries of uranium and molybdenum from synthetic samples ranged from 96.7 to 113.4% for uranium and from 96.7 to 110.5% for molybdate for a range of recommended conditions.  相似文献   

16.
Microwave plasma torch (MPT), traditionally used as the light source for atomic emission spectrophotometry, has been employed as the ambient ionization source for sensitive detection of uranium in various ground water samples with widely available ion trap mass spectrometer. In the full‐scan mass spectra obtained in the negative ion detection mode, uranium signal was featured by the uranyl nitrate complexes (e.g. [UO2(NO3)3]?), which yielded characteristic fragments in the tandem mass spectrometry experiments, allowing confident detection of trace uranium in water samples without sample pretreatment. Under the optimal experimental conditions, the calibration curves were linearly responded within the concentration levels ranged in 10–1000 µg·l?1, with the limit of detection (LOD) of 31.03 ng·l?1. The relative standard deviations (RSD) values were 2.1–5.8% for the given samples at 100 µg·l?1. The newly established method has been applied to direct detection of uranium in practical mine water samples, providing reasonable recoveries 90.94–112.36% for all the samples tested. The analysis of a single sample was completed within 30 s, showing a promising potential of the method for sensitive detection of trace uranium with improved throughput. Copyright © 2016 John Wiley & Sons, Ltd.  相似文献   

17.
The interaction of uranium(VI) with carbonate ions was studied with absorption spectroscopy and time-resolved laser-induced fluorescence spectroscopy due to the importance of these complexes in environmental relevant waters. In the pH range from 2 to 11 the influence of the temperature on the spectra was studied to check changes in the abundances of several binding forms. It was found that several binding forms are predominant at different temperatures and pH values. This observation can be explained by speciation changes due to the dependence of chemical equilibria on the temperature. Furthermore photoluminescence spectra of aqueous solutions of uranyl carbonate complexes were observed at ambient temperatures for the first time and single component absorption spectra of the uranyl carbonate complexes UO2(CO3)3 4− and UO2(CO3)2 2− were derived.  相似文献   

18.
Room-temperature vibrational and photoluminescence (PL) spectra of a natural, rare hydrated calcium copper uranyl carbonate mineral, voglite (Ca2Cu(UO2)(CO3)4·6H2O) are recorded and discussed in details. Vibrational spectroscopy gives information about the structure of voglite, which is still missing due to its unknown crystallographic features. By comparison with other uranyl carbonates and sulfates, a strong Raman line occurring at 834 cm−1 is assigned to the ν1(UO2)2+ symmetric stretching vibration rather than to the ν2(CO3)2− out-of-plane bending vibration. The ν3(UO2)2+ antisymmetric stretching vibration is tentatively identified at 897 cm−1 from infrared (IR) spectroscopy. Several well resolved bands found at 1074,1092, 1381, 1566 cm−1 in the Raman and 1046, 1114, 1145, 1376, 1426, 1510, 1561 cm−1 in the IR are ascribed to symmetric and antisymmetric stretching motions of the carbonate units. The presence of all these intense vibrational bands points to different CO bond lengths. The infrared water band is well structured, suggesting a few different OH moieties in the crystal. Original micro-PL spectra show a manifold of vibronic features whose energy spacing is close to the frequency of the symmetric OUO stretching vibration and confirms the uranium origin of the most intense Raman band. The study suggests that voglite structure has no inversion centers, a low symmetry, and contains molecular units similar to those of the parent phases, andersonite or liebigite, like uranyl tricarbonate clusters (UTC). The existence of these UTCs in voglite is confirmed by density functional theory calculations. A new assignment of all vibrational modes is proposed.  相似文献   

19.
利用化学种态分析软件CHEMSPEC计算了低浓缩铀靶辐照后溶液中铀(U)的化学种态分布及其主要裂变元素对U化学种态的影响。结果表明,在单组分体系中,pH值和铀酰浓度都会显著影响U的化学种态分布。随着铀酰浓度的增大,溶液中将会生成多核配合物;在较高的NO3-浓度下,U在溶液中主要以UO22+和UO2NO3+的形式存在。CO2对不同浓度铀的种态分布影响结果表明,当铀酰浓度较低时,铀的化学种态多以碳酸铀酰的形式存在;当铀酰浓度较高时,铀的化学种态多以氢氧铀酰或柱铀矿沉淀的形式存在。计算发现,当裂片元素Tc、I、Mo的浓度小于0.01mol·L-1并分别以TcO4-、I-、MoO42-的种态存在时,这些裂片元素不改变铀的各化学种态的分布。  相似文献   

20.
Understanding of the bonding nature of uranyl and various ligands is the key for designing robust sequestering agents for uranium extraction from seawater. In this paper thermodynamic properties related to the complexation reaction of uranyl(VI) in aqueous solution (i.e. existing in the form of UO2(H2O)5 2+) by several typical ligands (L) including acetate (CH3CO2 ?), bicarbonate (HOCO2 ?), carbonate (CO3 2?), CH3(NH2)CNO? (acetamidoximate, AO?) and glutarimidedioximate (denoted as GDO2?) have been investigated by using relativistic density functional theory (DFT). The geometries, vibrational frequencies, natural net charges, and bond orders of the formed uranyl-L complexes in aqueous solution are studied. Based on the DFT analysis we show that the binding interaction between uranyl and amidoximate ligand is the strongest among the selected complexes. The thermodynamics of the complexation reaction are examined, and the calculated results show that complexation of uranyl with amidoximate ligands is most preferred thermodynamically. Besides, reaction paths of the substitution complexation of solvated uranyl by acetate and AO? have been studied, respectively. We have obtained two minima along the reaction path of solvated uranyl with acetate, the monodentate-acetate complex and the bidentate-acetate one, while only one minimum involving monodentate-AO complex has been located for AO? ligand. Comparing the energy barriers of the two reaction paths, we find that complexation of uranyl with AO? is more difficult in kinetics, though it is more preferable in thermodynamics. These results show that theoretical studies can help to select efficient ligands with fine-tuned thermodynamic and kinetic properties for binding uranyl in seawater.  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