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 250 毫秒
1.
The kinetics of U(IV) oxidation with atmospheric oxygen in NaHCO3 solutions was studied by spectrophotometry. In 1 M NaHCO3 at [U(IV)]0 = 20 mM, an induction period is observed, which virtually disappears with decreasing [U(IV)]0 to 1.0 mM. The induction period is caused by the fact that initially U(IV) exists in a weakly active polymeric form. Addition of U(VI) to the initial solution accelerates the oxidation. In a 1 M NaHCO3 solution containing 0.1–1.0 mM U(IV), the U(IV) loss follows the first-order rate law with respect to U(IV) and O2. The pseudo-first-order rate constants, bimolecular rate constants, and activation energy of the U(IV) oxidation were calculated. In dilute NaHCO3 solutions (0.5–0.01 M), the hydrolysis and polymerization of U(IV) become more pronounced. The autocatalysis mechanism presumably involves formation of a complex [U(IV) · U(VI)] with which O2 reacts faster than with U(IV). Oxidation of U(IV) occurs by the two-electron charge-transfer mechanism.  相似文献   

2.
containing U(IV) polymer start to form. With an increase in pH from 1.5 to 4 or in temperature, the induction period becomes shorter. Under anaerobic conditions, the colloidal solution is stable for more than a month. Centrifugation at 8000 rpm (5500g) allows separation of the colloidal particles from the liquid phase. The colloid slowly dissolves in mineral acids saturated with argon or in a K2CO3 solution, whereas precipitates of individual freshly prepared U(IV) and U(VI) hydroxides dissolve rapidly. Short UV irradia-of a UO2(ClO4)2 solution saturated with argon and containing ethanol (pH 2.5) results in the appearance of U(V) which then disproportionates, and U(IV) forms with U(VI) a black colloid similar to that arising on mixing U(IV) and U(VI) solutions.  相似文献   

3.
Decomposition of aqueous suspensions of uranium(IV) oxalate under the action of an ozone–oxygen mixture was studied. The process occurs in two steps. In the first step, the U(IV) oxidation with the formation of oxalic acid uranyl solutions prevails. The second step involves decomposition of oxalate ions and hydrolysis of uranyl ions. An increase in temperature accelerates the transformation of uranium(IV) oxalate into uranium(VI) hydroxide compounds. In solutions containing KBr or UO2Br2, the following reaction occurs: O3 + Br → O2 + BrO. The arising hypobromite ions and hypobromous acid oxidize uranium(IV) oxalate extremely efficiently. The possible mechanism of ozonation of aqueous uranium(IV) oxalate suspensions is discussed.  相似文献   

4.
The kinetics of U(VI) accumulation in the phase of U(IV) hydroxide and of Np(V) in the phase of neptunium(IV) hydroxide, and also the solubility of the formed mixed-valence U(IV)-U(IV) and Np(IV)-Np(V) hydroxides in simulated groundwater (SGW, pH 8.5) and 0.1 M NaClO4 (pH 6.9) solutions was studied. It was found that the structure of the mixed U(IV–VI) hydroxide obtained by both oxidation of U(IV) hydroxide with atmospheric oxygen and alkaline precipitation from aqueous solution containing simultaneously U(IV) and U(VI) did not affect its solubility at the U(VI) content in the system exceeding 16%. The solubility of mixed-valence U(IV–VI) hydroxides in SGW and 0.1 M NaClO4 is (3.6±1.9) × 10?4 and (4.3 ± 1.7) × 10?4 M, respectively. The mixed Np(IV–V) hydroxide containing from 8 to 90% Np(V) has a peculiar structure controlling its properties. The solubility of the mixed-valence Np(IV–V) hydroxide in SGW [(6.5 ± 1.5) × 10?6 M] and 0.1 M NaClO4 [(6.1±2.4) × 10?6 M] is virtually equal. Its solubility is about three orders of magnitude as high as that of pure Np(OH)4 (10?9–10?8 M), but considerably smaller than that of NpO2(OH) (~7 × 10?4 M). The solubility is independent of the preparation procedure [oxidation of Np(OH)4 with atmospheric oxygen or precipitation from Np(IV) + Np(V) solutions]. The solubility of the mixed-valence Np hydroxide does not increase and even somewhat decreases [to (1.4±0.7) × 10?6 M] in the course of prolonged storage (for more than a year).  相似文献   

5.
The reaction of the ozone–oxygen mixture with aqueous suspensions of Np(IV) and Pu(IV) oxalates was studied. Both metal cations and oxalate anions are oxidized in the process. The final products are Np(VI) and Pu(VI) hydroxides. The composition of Np(VI) hydroxide was confirmed by X-ray diffraction analysis. Oxidation of Np(IV) oxalate with oxygen leads to the accumulation of Np(V) oxalate and oxalic acid in the solution. At incomplete oxidation of Np(IV) oxalate with ozone in water, Np(V) is also accumulated. Heating considerably accelerates the ozonation. The possible reaction mechanism is briefly discussed. The Np(V) and Np(VI) ions participate in the catalytic cycle of the decomposition of oxalate ions with ozone.  相似文献   

6.
Extraction of trace amounts of Pu(IV), U(VI), Th(IV), Am(III), and REEs(III) from nitric acid solutions with o-methyleneoxyphenyldiphosphine dioxides with phenyl (I) and butyl (II) substituents at the phosphoryl group is studied. The stoichiometry of extractable complexes in dichloroethane is determined. Changing the butyl substituents for phenyl ones increases the effective extraction constants of Pu(IV), Th(IV), and Am(III) and decreases that of U(VI). In extraction from 3 M HNO3, the Pu(IV)/U(VI) and Th(IV)/U(VI) partition coefficients are higher with reagent I, and the Pu(IV)/Am(III) and U(VI)/Am(III) partition coefficients, with reagent II.  相似文献   

7.
The complexation of U(VI), Np(VI), and Pu(VI) and of Np(V) with 1,2,3- and 1,2,4-benzenetricarboxylic acids (BTC) in aqueous solutions was studied in wide ranges of pH and actinyl ion concentrations. The compositions of the forming hexavalent actinide complexes were determined. Their apparent stability constants β1′ depend on pH of the solution: in the pH range 2–4, logβ1′ from 2 to 4 for the complexes of U(VI), Np(VI), and Pu(VI) with 1,2,3-BTC and from 1.5 to 3.5 for the complexes with 1,2,4-BTC. For Np(V), the β1′ values are close with both acids, and at equal pH values the Np(V) complexes are less stable than the An(VI) complexes (An = U, Np, Pu). With an increase in pH from ~3 to 6.2–6.9, logβ1′ of the Np(V) complexes increases approximately from 0.5 to 3. Solid U(VI) complexes with 1,2,3- and 1,2,4-benzenetricarboxylic acids were synthesized by the hydrothermal method, their crystal structure was determined, and the IR spectra were examined.  相似文献   

8.
The behavior of Pu(VI) and Pu(V) in CH3COOH (HAc)-H2O solutions was studied by spectrophotometry. The absorption spectrum of Pu(VI) does not change on adding HAc to a concentration of 5 M in the presence of 0.5–1.0 M HClO4, but in solutions containing less than 0.001 M mineral acid, changes in the spectrum are observed at HAc concentration of 0.6 M.he major absorption band of PuO 2 2+ ions, caused by an f-f transition, with increasing [HAc] is shifted from 830.6 to 836 nm, with a simultaneous decrease in the absorption intensity, which is due to formation of 1: 1 complexes of Pu(VI) with Ac? ions. In anhydrous HAc, the peak intensity increases again, owing to total change in the composition of the solvation shell. Pu(V) is unstable in 1–17 M HAc solutions and disproportionates to form Pu(VI) and Pu(IV). The Pu(V) loss follows a second-order rate law with respect to [Pu(V)] and accelerates with increasing HAc concentration. The reaction products exert opposite effects on the reaction rate: Pu(IV) accelerates the consumption of Pu(V), whereas Pu(VI) does not affect the process in dilute HAc solutions but decelerates the disproportionation in concentrated solutions owing to formation of a cation-cation complex with Pu(V).  相似文献   

9.
Reactions of Pu(IV) and Np(VI) with organic reducing agents of various types (substituted hydroxylamines, oximes, aldehydes, etc.) in tributyl phosphate solutions containing nitric acid were studied spectrophotometrically. The molar extinction coefficients of neptunium and plutonium in various oxidation states [Np(IV,V,VI), Pu(III,IV,VI)] in TBP solutions were determined as influenced by HNO3 and H2O concentrations and temperature. It was found that organic reducing agents at low HNO3 concentration convert plutonium and neptunium to Pu(III) and Np(V), respectively. With increasing HNO3 concentration Pu(III) and Np(V) are partly oxidized back to Pu(IV) and Np(VI), respectively, by reaction with nitrous acid. The rate constants of Pu(VI) and Np(VI) reduction and Np(V) oxidation as influenced by concentration of organic reducing agents and HNO3 were evaluted from the kinetic data.  相似文献   

10.
Photochemical reactions of Np(VI), Np(V), and Np(IV) in aqueous solutions containing HCO3 - and CO3 2- anions were studied spectrophotometrically. Two parallel photoreactions, oxidation of Np(V) to Np(VI) and reduction of Np(VI) to Np(V), were revealed. The quantum efficiency of Np(VI) photoreduction in 1.95 M Na2CO3 is 0.003. As the pH of the solution decreases, Np(VI) photoreduction is decelerated and Np(V) photooxidation is accelerated. The photoconversion of Np(V) into Np(VI) in 1 M NaHCO3 is 94-95%. The presence of oxygen has no effect on the oxidation. Neptunium(IV) is oxidized to Np(V) and Np(VI).  相似文献   

11.
Sorption of Cs, Pu(IV), Np(V), and U(VI) with bentonite from solutions was studied. Physicochemical species of radionuclides in the liquid phase were determined, the sorption mechanisms were established, and the influence of bentonite colloids on the behavior of radionuclides was studied. It was shown that Cs is sorbed by the ion-exchange mechanism, whereas the sorption of actinides at pH > 5 is governed by the reaction with surface hydroxy groups of betonite, and at pH < 5 the competing processes are ion exchange and complex formation. Reduction of Np(V) and U(VI) to Np(IV) and U(IV) in the solution with Fe(II) compounds present in the system was proved by the extraction method. Various methods of separating the solid phase were used in studying the dependence of the distribution coefficients of Np and Pu on the ratio of pore water and bentonite; it was shown that Np and Pu are sorbed on bentonite colloids.  相似文献   

12.
The behavior of Pu(VI), Pu(V), and Pu(IV) in the HCOOH-H2O system was studied by spectrophotometry. The Pu(VI) absorption spectrum in solutions containing less than 1 mM HClO4 changes on adding HCOOH to a concentration of 0.53 M. Along with a decrease in the intensity of the absorption maximum at 830.6 nm, corresponding to an f-f transition in the Pu22+ aqua ion, a new band arises with the maximum shifted to 834.5 nm. These transformations are due to formation of a Pu(VI) formate complex (1: 1). The Pu(IV) absorption spectra in HCOOH solutions vary insignificantly in going from 3.0 to 9.0 M HCOOH and are similar to the spectrum of Pu(IV) in a 0.88 M HCOOH + 0.41 M NaHCOO + 0.88 M NaClO4 solution, which indicates that the composition of the Pu(IV) formate complexes is constant. Pu(V) is unstable in HCOOH solutions and disproportionates to form Pu(VI) and Pu(IV). The reaction rate is approximately proportional to [Pu(V)]2 and grows with an increase in [HCOOH]. The reaction products affect the reaction rate: Pu(IV) accelerates the process, and Pu(VI) decelerates the consumption of Pu(V) by binding Pu(V) in a cationcation complex. The disproportionation occurs via formation of a Pu(V)-Pu(V) cation-cation complex whose thermal excitation yields an activated complex with its subsequent decomposition to Pu(VI) and Pu(IV).  相似文献   

13.
The reduction of Pu(IV) and Np(VI) with carbohydrazide (NH2NH)2CO in 1–6 M HNO3 solutions was studied. The Pu(IV) reduction is described by a first-order rate equation with respect to Pu(IV). At [HNO3] ≥ 3 M, the reaction becomes reversible. The rate constants of the forward and reverse reactions were determined, and their activation energies were estimated. Neptunium(VI) is reduced to Np(V) at a high rate, whereas the subsequent reduction of Np(V) to Np(IV) is considerably slower and is catalyzed by Fe and Tc ions. The possibility of using carbohydrazide for stabilizing desired combinations of Pu and Np valence states was examined.  相似文献   

14.
The mechanisms of photo-catalytic reduction of Cr(VI) were investigated in acidic solutions with and without Fe(III). In a system without Fe(III), no Cr(VI) reduction was observed in dark conditions; conversely, under light conditions, the reduction reaction rate increased to 0.011 and 0.020microM min(-1) at pH 2 and pH 1, respectively, indicating the occurrence of Cr(VI) photo-reduction. The Cr(VI) photo-reduction reaction was induced by the photolysis of water molecules, leading to O(2) production. Upon the addition of Fe(III), the photo-reduction rate of Cr(VI) was significantly enhanced due to the formation of Fe(II), which is the photolytic product of FeCl(2)(+) and the electron donor for Cr(VI) reduction. However, with the same concentration of FeCl complexes, a strong inhibition of Cr(VI) reduction at pH 2 was observed, compared with pH 1. A possible explanation is that FeOH(2+) becomes predominant with increasing pH and that its photolytic product, the OH free radical, is an oxidant for Fe(II) and Cr(III) and can compromise Cr(VI) reduction. The kinetic result of each photo-reduction reaction pathway shows zero-order kinetics, suggesting that the photolysis reaction of H(2)O or FeCl(2+) is the rate-determining step in each pathway. The results also show the potential of developing a homogeneous photo-catalytic method to treat Cr(VI)-containing water.  相似文献   

15.
The isotherm of uranyl nitrate extraction with TBP in diluent without salting-out agents with the formation of the disolvate is described by a chemical equation with the concentration constant of 10 without correction coefficients in a wide range of component concentrations. The concentration of “free” TBP in the dependence of the extraction of U(VI) microamounts on the acid concentration was calculated using this value, and it was shown that the formal stoichiometric ratio TBP: HNO3 in the extract is 0.75. To describe this effect, a set of reactions was suggested, taking into account the role of extracted water, and the corresponding concentration equilibrium constants were calculated without corrections. The values obtained differ significantly from those suggested previously. The constants obtained adequately describe the joint extraction of U(VI) and HNO3 at the acid concentration of 0.3–8 M and extract loading with U(VI) of up to 85% of its capacity. At low extract loadings and acid concentration lower than 1 M, weak hydrolysis of uranyl ion and extraction of U(VI) hydrate solvate were taken into account. For the extraction of Np(VI) and Pu(VI) with TBP in diluent, the extraction constants that do not vary with the U(VI) and HNO3 concentrations were calculated taking into account the formation of the Np(VI) and Pu(VI) hydrate solvates.  相似文献   

16.
Sludge, soil and leachate samples collected from a chromium-contaminated tannery waste dumping site in Kanpur, India, were found to contain considerable amounts of Cr(VI), despite the fresh tannery sludge containing little or no Cr(VI). Literature reports suggested that dry Cr(III) precipitates could be converted to Cr(VI) when heated in the presence of oxygen. Also, Cr(III) in aqueous phase could be oxidized through interaction with manganese dioxide (MnO2) surface to Cr(VI). Measurement of manganese in the sludge samples collected from the site showed concentrations up to 0.6 mg/g. Based on equilibrium calculations, it was determined that both dry phase Cr(III) oxidation by atmospheric oxygen and aqueous phase Cr(III) oxidation by MnO2 surface were thermodynamically feasible. It was further suggested that in aqueous phase, manganese may act effectively as an electron transporter between Cr(III) and dissolved oxygen during Cr(III) oxidation, leading to regeneration of MnO2 solid phase. Further, as dissolved Cr(III) is oxidized, dissolution of Cr(OH3) will take place to maintain the equilibrium between the dissolved and solid phases of Cr(III). In the pH range of 3-10, and at oxygen partial pressure (P(O2)) of 10(-6) atm or higher, equilibrium conditions stipulate nearly complete conversion of Cr(III) to Cr(VI). At P(O2) of 10(-20) atm or lower, very little Cr(VI) is expected to be present under equilibrium conditions. In the intermediate P(O2) regions, incomplete dissolution of the Cr(OH3) solid phase and only partial conversion of chromium from +3 to the +6 oxidation state is expected, especially at lower pH values.  相似文献   

17.
The possibilities of using 2,2′-bipyridine-6,6′-dicarboxamides in modified versions of the UNEX process were examined. The stability constants of complexes of Eu(III), Am(III), Th(IV), U(VI), and Np(V) ions with certain diamides substituted in the amide moiety or pyridine ring in acetonitrile were determined. The logarithms of the stability constants exceeded 7 for Eu(III) and reached 9 for 5f elements. The ability of the selected diamides in nitrobenzene and trifluoromethyl phenyl sulfone (FS-13) diluents to extract Am(III), Th(IV), U(VI), and Eu(III) was studied.  相似文献   

18.
The stability constants of the complexes of U(VI), Np(VI), and Pu(VI) with the heteropolyanions (HPAs) P2W17O 61 10? , SiW11O 39 8? , and PW11O 39 7? in solutions with pH from ?0.3 (2 M H+) to 5–5.5 in the presence of Na or K salts (up to 2 M) and without them were measured. All the complexes have exclusively the 1: 1 composition; their stability constants βM(VI) in neutral solutions at a low ionic strength are close to 108 1 mol?1. In 0.1–2.0 M acid solutions, log βM(VI) for the complexes with P2W17O 61 10? is within 1.4–3.9. The slope of the pH dependence of log βM(VI) does not exceed 1.75; this fact suggests that no more than two protons are displaced from HPA upon complexation in acid solutions. In the presence of 1–2 M sodium salts, the βM(VI) values reach a maximum at pH ~3 and drastically decrease with a further increase in pH. Actinides(VI) interact with HPAs appreciably more weakly than do actinides(III), which is apparently due to the fact that the denticity of HPAs in the complexes with An(VI), apparently, does not exceed 2.  相似文献   

19.
Extraction of microamounts of REE(III), U(VI), and Th(IV) with solutions of carbamoylmethylphosphine oxides (CMPOs) in organic diluents from aqueous HNO3 solutions containing lithium bis[(trifluoromethyl) sulfonyl]imide (LiTf2N) was studied. The efficiency of the REE(III), U(VI), and Th(IV) extraction from nitric acid solutions with CMPO solutions considerably increases in the presence of Tf2N ions in the aqueous phase. The stoichiometry of the extractable complexes was determined, and the influence of the structure of the CMPO molecule, kind of organic diluent, and aqueous phase composition on the efficiency of the U(VI), Th(IV), and REE(III) extraction into the organic phase was considered.  相似文献   

20.
Treatment of uranium-containing waters by the complexation/ultrafiltration (COUF) method using polyethylenimine (PEI) and disodium dihydrogen ethylenediaminetetraacetate (EDTA) as complexing agents was studied. The influence of pH of the solutions being purified on the treatment process was examined. With PEI and PEI-EDTA mixture at pH 7.0 and higher, the maximum possible coefficient of U(VI) retention by a UPM-20 membrane (0.999) is attained. Addition of EDTA to a solution containing a mixture of U and PEI improves the U(VI) retention, especially in the acidic pH region, owing to formation of mixed U(VI)-EDTA-PEI complexes. The formation constant of this complex was calculated by mathematical simulation.  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