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.
Results of oxidation experiments on high-burn-up UO2 are presented where fission-product vaporisation and release rates have been measured by on-line mass spectrometry as a function of time/temperature during thermal annealing treatments in a Knudsen cell under controlled oxygen atmosphere. Fractional release curves of fission gas and other less volatile fission products in the temperature range 800-2000 K were obtained from BWR fuel samples of 65 GWd t−1 burn-up and oxidized to U3O8 at low temperature. The diffusion enthalpy of gaseous fission products and helium in different structures of U3O8 was determined.  相似文献   

2.
The reduction of U3O8 pellets to UO2+x has been investigated at 1300 °C in H2, Ar and CO2 gas atmospheres by TGA, SEM, and X-ray diffraction. The selected U3O8 pellet was prepared by sintering a U3O8 powder compact. The TGA results show that the reduction rate is fastest in H2 gas, and X-ray diffraction indicates that U3O8 reduces to UO2+x without any intermediate phase. The reduced pellet, UO2+x, has a special grain structure that consists of equiaxed grains at the surface, columnar grains in the middle, and equiaxed grains in the center. The equiaxed grains and columnar grains are much smaller in H2 gas than in Ar or CO2 gas. The reducing gases significantly influence the morphology of the grain structure. This difference can be explained in terms of a relation between oxygen potential and critical nucleus size during the reduction.  相似文献   

3.
Solid state reactions of UO2 and ZrO2 in mild oxidizing condition followed by reduction at 1673 K showed enhanced solubility up to 35 mol% of zirconium in UO2 forming cubic fluorite type ZryU1−yO2 solid solution. The lattice parameters and O/M (M = U + Zr) ratios of the solid solutions, ZryU1−yO2+x, prepared in different gas streams were investigated. The lattice parameters of these solid solutions were expressed as a linear equation of x and y: a0 (nm) = 0.54704 − 0.021x - 0.030y. The oxidation of these solid solutions for 0.1 ? y ? 0.2 resulted in cubic phase MO2+x up to700 K and single orthorhombic zirconium substituted α-U3O8 phase at 1000 K. The kinetics of oxidation of ZryU1−yO2 in air for y = 0-0.35 were also studied using thermogravimetry. The specific heat capacities of ZryU1−yO2 (y = 0-0.35) were measured using heat flux differential scanning calorimetry in the temperature range of 334-860 K.  相似文献   

4.
The thermal diffusivities of UO2 and U4O9 were measured by the laser flash method at temperatures ranging from 100 to 300 K. The phonon mean free path and the thermal conductivity were calculated from the obtained thermal diffusivity data and the heat capacity. The structure of the u4o9 is closely related to the UO2 structure with an excess oxygen atom per unit cell in U4O9. As the excess oxygen atoms increase the anharmonicity of the lattice vibration, the phonon mean free path in U4O9 decreases. Therefore, the thermal conductivity of U4O9 is much lower than that of UO2 and increases slightly with increasing temperature due to the rise in heat capacity.  相似文献   

5.
We have performed ab initio total energy calculations to investigate the behavior of helium and its diffusion properties in uranium dioxide (UO2). Our investigations are based on the density functional theory within the generalized gradient approximation (GGA). The trapping behavior of He in UO2 has been modeled with a supercell containing 96-atoms as well as uranium and oxygen vacancy trapping sites. The calculated incorporation energies show that for He a uranium vacancy is more stable than an oxygen vacancy or an octahedral interstitial site (OIS). Interstitial site hopping is found to be the rate-determining mechanism of the He diffusion process and the corresponding migration energy is computed as 2.79 eV at 0 K (with the spin-orbit coupling (SOC) included), and as 2.09 eV by using the thermally expanded lattice parameter of UO2 at 1200 K, which is relatively close to the experimental value of 2.0 eV. The lattice expansion coefficient of He-induced swelling of UO2 is calculated as 9 × 10−2. For two He atoms, we have found that they form a dumbbell configuration if they are close enough to each other, and that the lattice expansion induced by a dumbbell is larger than by two distant interstitial He atoms. The clustering tendency of He has been studied for small clusters of up to six He atoms. We find that He strongly tends to cluster in the vicinity of an OIS, and that the collective action of the He atoms is sufficient to spontaneously create additional point defects around the He cluster in the UO2 lattice.  相似文献   

6.
The addition of Th to U-based fuels increases resistance to corrosion due to differences in redox-chemistry and electronic properties between UO2 and ThO2. Quantum-mechanical techniques were used to calculate surface energy trends for ThO2, resulting in (1 1 1) < (1 1 0) < (1 0 0). Adsorption energy trends were calculated for water and oxygen on the stable (1 1 1) surface of UO2 and ThO2, and the effect of model set-up on these trends was evaluated. Molecular water is more stable than dissociated water on both binary oxides. Oxidation rates for atomic oxygen interacting with defect-free UO2(1 1 1) were calculated to be extremely slow if no water is present, but nearly instantaneous if water is present. The semi-conducting nature of UO2 is found to enhance the adsorption of oxygen in the presence of water through changes in near-surface electronic structure; the same effect is not observed on the insulating surface of ThO2.  相似文献   

7.
Oxygen potentials of hypo-stoichiometric Lu-doped UO2, (U0.80Lu0.20)O2−x, were experimentally investigated by thermogravimetric analysis using H2O/H2 gas equilibria at 1173, 1273 and 1473 K. The oxygen potentials of (U,Lu)O2−x were higher than those of other forms of rare earth-doped UO2, specifically (U,Nd)O2−x, (U,Gd)O2−x, and (U,Er)O2−x. Slope analyses for plots of oxygen potential versus deviation from stoichiometry indicated that (U0.80Lu0.20)O2−x had a similar defect structure to that of the other forms of rare earth-doped UO2. A relationship between the effective ionic radii and oxygen potentials was found for the hypo-stoichiometric rare earth-doped UO2.  相似文献   

8.
Erbium is considered as a slow burnable poison suitable for use in light water reactors (LWRs). Addition of a small amount of Er2O3 to all UO2 pellets will make it possible to develop super high burnup fuels in Japanese nuclear facilities which are now under the restriction of the upper limit of 235U enrichment. When utilizing the (U,Er)O2 fuels, it is very important to understand the thermal and mechanical properties. Here we show the characterization results of (U1−xErx)O2 (0 ? x ? 0.1). We measured their thermal and mechanical properties and investigated the effect of Er addition on these properties of (U,Er)O2. All Er completely dissolved in UO2, and the lattice parameter decreased linearly with the Er content. Both the thermal conductivity and Young’s modulus of (U,Er)O2 decreased with the Er content. These results would be useful for us in evaluating the performance of the (U,Er)O2 fuels in LWRs.  相似文献   

9.
Stoichiometries in (U0.7Pu0.3)Ox and (U0.8Pu0.2)Ox were analyzed with the experimental data of oxygen potential based on point defect chemistry. The relationship between the deviation x of stoichiometric composition and the oxygen partial pressure PO2 was evaluated using a Kröger-Vink diagram. The concentrations of the point defects in uranium and plutonium mixed oxide (MOX) were estimated from the measurement data of oxygen potentials as functions of temperature and PO2. The analysis results showed that x was proportional to near the stoichiometric region of both (U0.7Pu0.3)Ox and (U0.8Pu0.2)Ox, which suggested that intrinsic ionization was the dominant defect. A model to calculate oxygen potential was derived and it represented the experimental data accurately. Further, the model estimated the thermodynamic data, and , of stoichiometric (U0.7Pu0.3)O2.00 and (U0.8Pu0.2)O2.00 as −552.5 kJ·mol−1 and −149.7 J·mol−1, and −674.0 kJ · mol-1 and −219.4 J · mol−1, respectively.  相似文献   

10.
The solubility of uranium dioxide (UO2) was measured in real and synthetic Boom Clay waters with varying concentrations of humic acids and carbonate under reducing conditions at 20 °C. Uranium concentrations in function of time suggest the reduction of U(VI) to U(IV) by the humic acids which is occurring faster in real clay water than in synthetic clay waters. Humic acids induce also a competition to complex U(VI) in carbonate-containing solution, but they are not able to control the uranium concentration at high bicarbonate concentration (0.02 mol dm−3). Nevertheless they may play a role at low carbonate concentration. In our experimental conditions, the geochemical calculations indicate that two uranium secondary phases (U4O9 and UO2(c)) are susceptible to control the uranium concentration in solution. These calculations are in good agreement with results of the X-ray photoelectron spectroscopy. At the end of tests, uranium concentrations reach steady-state values between 3 × 10−8 and 5 × 10−8 mol dm−3 in the bicarbonate-rich solutions. Although these concentrations are considered as conservative, they are 10-100 times higher than in natural Boom Clay. The consequence is that spent fuel could slowly dissolve in the interstitial clay water undersaturated with respect to UO2/UO2+x of the fuel.  相似文献   

11.
Thermal diffusivities of UO2 and (U, Gd)O2 pellets irradiated in a commercial reactor (maximum burnups: 60 GWd/t for UO2 and 50 GWd/t for (U, Gd)O2) were measured up to about 2000 K by using a laser flash method. The thermal diffusivities of irradiated UO2 and (U, Gd)O2 pellets showed hysteresis phenomena: the thermal diffusivities of irradiated pellets began to recover above 750 K and almost completely recovered after annealing above 1400 K. The thermal diffusivities after recovery were close to those of simulated soluble fission products (FPs)-doped UO2 and (U, Gd)O2 pellets, which corresponded with the recovery behaviors of irradiation defects for UO2 and (U, Gd)O2 pellets. The thermal conductivities for irradiated UO2 and (U, Gd)O2 pellets were evaluated from measured thermal diffusivities, specific heat capacities of unirradiated UO2 pellets and measured sample densities. The difference in relative thermal conductivities between irradiated UO2 and (U, Gd)O2 pellets tended to become insignificant with increasing burnups of samples.  相似文献   

12.
The effects of a powder treatment, the sintering temperature and the sintering time on the grain growth of UO2 pellets were investigated in air to obtain UO2 pellets with large grains. Air could be used for sintering because an oxidation path above 1803 K does not pass through a two-phase (UO2+x + U3O8−z) region. The UO2 pellets sintered by the CO2-air-CO2-H2 process consisted of a single grain or some large grains in the order of several millimeters.  相似文献   

13.
Solid-state chemical investigations have established that in the compositional range UO2-UO2.67-ThO3 of the U-Th-O ternary system, the following single-phase domains exist: U3O8, which does not dissolve any ThO2 in the solid state; an ordered M4O9 phase on the section between U4O9 and U2Th2O9, below ≈ 1150 °C; and a phase with fluorite structure which occupies a large part of the system and which at 1250 °C is bounded by the compositions UO2-UO2.25 (U0.43, ThO0.57)O0.25-ThO3. The maximum O/M ratio of the “fluorite” phase is O:(U + Th) = 2.25. The highest oxidation valency of uranium is 5.30; this value falls as more thorium oxide is incorporated in the (U.Th)O2 + x “fluorite” phase.  相似文献   

14.
The thermal conductivity, Young’s modulus, and hardness of (U0.65−xCe0.3Pr0.05Ndx)O2 (x = 0.01, 0.08, 0.12) were evaluated and the effect of Pr and Nd addition on the properties of (U, Ce)O2 were studied. The polycrystalline high-density pellets were prepared with solid state reactions of UO2, CeO2, Pr2O3, and Nd2O3. We confirmed that all Ce, Pr, and Nd dissolved in UO2 and formed solid solutions of (U, Ce, Pr, Nd)O2. We revealed that the thermal conductivity of (U0.65−xCe0.3Pr0.05Ndx)O2 (x = 0.12) was up to 25% lower than that of x = 0.01 at room temperature. The Young’s modulus of (U0.65−xCe0.3Pr0.05Ndx)O2 decreased with x, whereas the hardness values were constant in the investigated x range.  相似文献   

15.
Heat capacities and enthalpy increments of solid solutions Th1−yUyO2(s) (y = 0.0196, 0.0392, 0.0588, 0.098, 0.1964) and Simfuel (y = 0.0196) were measured by using a differential scanning calorimeter and a high temperature drop calorimeter. The heat capacities were measured in two temperature ranges: 127-305 K and 305-845 K and enthalpy increments were determined in the temperature range 891-1698 K. A heat capacity expression as a function of uranium content y and temperature and a set of self-consistent thermodynamic functions for Th1−yUyO2(s) were computed from present work and the literature data. The oxygen potentials of Th1−yUyO2+x(s) have been calculated and expressed as a polynomial functions of uranium content y, excess oxygen x and temperature T. The phase diagram, oxygen potential diagram of thorium-uranium-oxygen system and major vapour species over urania thoria mixed oxide have been computed using FactSage code.  相似文献   

16.
The effects of alpha dose-rate on UO2 dissolution were investigated by performing dissolution experiments with 238Pu-doped UO2 materials containing nominal alpha-activity levels of ∼1-100 Ci/kg UO2 (actual levels 0.4-80 Ci/kg UO2), in 0.1 M NaClO4 and in 0.1 M NaClO4 + 0.1 M carbonate. Dissolution rates increased less than 10-fold for an almost 100-fold increase in doping level and fall within the range of predictions of the Mixed Potential Model (a detailed mechanistic model for used fuel dissolution). Dissolution rates were lower in carbonate-free solutions and enrichment of 238Pu on the UO2 surface was suggested in carbonate solutions. Effective G values, defined as the ratio of the total amount of U dissolved divided by the maximum possible amount of U dissolved by radiolytically produced H2O2, increased with decreasing doping levels. This suggests that the dissolution reaction at high dose rates is limited by the reaction rate between UO2 and H2O2, but becomes increasingly limited by the rate of production of H2O2 at lower dose rates.  相似文献   

17.
In this work, we have studied the impact of Y2O3 on the kinetics of oxidative dissolution of UO2 and the consumption of H2O2. The second order kinetics of catalytic consumption of H2O2 on Y2O3 was investigated in aqueous Y2O3 powder suspensions by varying the solid surface area to solution volume ratio. The resulting second order rate constant is 10−8 m s−1, which is of the same magnitude as for the reaction between H2O2 and UO2. Powder experiments with mixtures of UO2 and Y2O3 show that Y2O3 has no effect on the oxidative dissolution of UO2, whereas the consumption of H2O2 seems to be slightly slower in the presence of Y2O3 and H2 respectively. UO2 pellets with solid inclusions of Y2O3 show a decrease in oxidative dissolution by a factor of 3.3 and 5.3 under inert and hydrogen atmosphere, respectively. The rate of H2O2 consumption is similar for all cases and is well in line with kinetic data from powder experiments. The effects of H2 and Y2O3 on the oxidative dissolution of UO2 under gamma irradiation are similar to those found in experiments with H2O2. No significant difference in dissolution between inert and reducing atmosphere can be observed for pure UO2.  相似文献   

18.
Uranium carbide dispersed in graphite was produced under vacuum by means of carbothermic reduction of different uranium oxides (UO2, U3O8 and UO3), using graphite as the source of carbon. The thermal process was monitored by mass spectrometry and the gas evolution confirmed the reduction of the U3O8 and UO3 oxides to UO2 before the carbothermic reaction, that started to occur at T > 1000 °C. XRD analysis confirmed the formation of α-UC2 and of a minor amount of UC. The morphology of the produced uranium carbide was not affected by the oxides employed as the source of uranium.  相似文献   

19.
The reactivity of H2 towards UO22+ has been studied experimentally using a PEEK coated autoclave where the UO22+ concentration in aqueous solution containing 2 mM carbonate was measured as a function of time at pH2∼40 bar. The experiments were performed in the temperature interval 74-100 °C. In addition, the suggested catalytic activity of UO2 on the reduction of UO22+ by H2 was investigated. The results clearly show that H2 is capable of reducing UO22+ to UO2 without the presence of a catalyst. The reaction is of first order with respect to UO22+. The activation energy for the process is 130 ± 24 kJ mol−1 and the rate constant is k298K=3.6×10−9 l mol−1 s−1. The activation enthalpy and entropy for the process was determined to 126 kJ mol−1 and 16.5 J mol−1 K−1, respectively. Traces of oxygen were shown to inhibit the reduction process. Hence, the suggested catalytic activity of freshly precipitated UO2 on the reduction of UO22+ by H2 could not be confirmed.  相似文献   

20.
Diffuse reflectance measurements were made over the wavenumber range of 4000-20,000 cm−1 at room temperature on monoclinic and stabilised ZrO2, together with Y2Ti2O7 having the pyrochlore structure, all of which were doped with U and sintered in various atmospheres. X-ray photoelectron spectroscopy measurements were also carried out on selected samples. In monoclinic and stabilised zirconia, U exhibited valence states of +4 and/or +5, depending on the sintering atmosphere and the presence of appropriate charge compensators. Using both diffuse reflectance and X-ray photoelectron spectroscopy, U was also observed as mainly U4+ and/or U5+ in U-doped Y2Ti2O7 sintered at 1400 °C in air or Ar, although a small amount of U6+ also appeared to be present in some U-doped Y2Ti2O7 samples heated in air.  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