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 0 毫秒
1.
The influence of a number of powder characteristics on the sintering behaviour of UO2 has been evaluated. A good correlation is found between the green and sintered densities (dg and ds) and the specific surface area as determined by Knudsen flow permeametry. This correlation, however, does not reveal the differences in behaviour which are borne out by the sintering index defined as: Is = (ds ? dg)/(100 ? dg). These differences are explained in terms of the absence or presence of open porosity within the powders, which can be measured by means of mercury porosimetry. Thermal treatments have a more pronounced effect on the sinterability of powders with open porosity than on those without.  相似文献   

2.
3.
We perform first-principles calculations of electronic structure and optical properties for UO2 and PuO2 based on the density functional theory using the generalized gradient approximation (GGA) + U scheme. The main features in orbital-resolved partial density of states for occupied f and p orbitals, unoccupied d orbitals, and related gaps are well reproduced compared to experimental observations. Based on the satisfactory ground-state electronic structure calculations, the dynamical dielectric function and related optical spectra, i.e., the reflectivity, adsorption coefficient, energy-loss, and refractive index spectrum, are obtained. These results are consistent with the available experiments.  相似文献   

4.
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   

5.
The kinetics of initial stage sintering of UO2 powder were reinvestigated, using Ar-10% H2 atmosphere. The effect of the addition of neodynium oxide was studied. The results revealed that surface and grain boundary diffusion mechanisms act simultaneously. The values of activation energies were found to be 48.48 ± 3.51 kcal/mole in the temperature range 870–942°C and 89.88 ± 9.87 kcal/mole in the temperature range of 942–1030°C for UO2, and 115.61 ± 7.77 kcal/mole in the temperature range 1030–1150°C for UO2 + Nd2O3. An important decrease in the calculated diffusion coefficient occurs by the addition of Nd2O3.  相似文献   

6.
7.
8.
9.
A new technique has been developed to study fission-induced densification and hot-pressing of UO2 at very low temperatures without complications from fracturing or other concurrent thermal effects. Thin disks of UO2 of 0.22 to 20% enrichment and differing microstructural stability, were irradiated in the core of the CP — reactor at temperatures below 200°C and at pressures from 1 atm to 20.7 MPa. Results indicate that the pressurizing medium, NaK, had penetrated the open porosity at high pressure and impeded densification. To rationalize this effect, the previously proposed models for radiationinduced densification are critically reviewed. Modifications to models involving thermal sintering and hot-pressing, and pore resolution appear the most tenable. The former mechanism leads to predictions that fission-induced hot-pressing can occur, and ex-reactor sintering and hot-pressing should correlate with in-reactor densification. The proximity of pores and grain boundaries is also stressed. The effect of NaK logging is rationalized by changes in pore surface energy and by stabilization of small pores aginst complete resolution.  相似文献   

10.
Ashby's method of constructing sintering diagrams has been modified to obtain relative contribution diagrams directly from the computer. It is endeavoured here to study the interplay of sintering variables and mechanisms and determine the factors that affect the participation of mechanisms in UO2. By studying the physical properties, it emerges that the order of inaccuracies is small in most cases and do not affect the diagrams. On the other hand, even a 10% error in activation energies, which is quite plausible, would make a significant difference to the diagram. The main criticism of Ashby's approach is that the numerous properties and equations used, communicate their inaccuracies to the diagrams and make them unreliable. Our study has considerably reduced the number of factors that need to be refined to make the sintering diagrams more meaningful.  相似文献   

11.
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   

12.
Using the multiphase equilibrium method for determination of wetting angles we determined the surface and grain boundary energies of polycrystalline stoichiometric UO2 at 1500, 1750 and 1900°C. The data of the surface energy of UO2 agree with most of the published results. Linear temperature functions were obtained by extrapolation for both quantities between 0 K and the melting point of UO2. Measurements of the interfacial energy in the UO2-Ni system yielded a linear dependence on temperature in the range investigated.  相似文献   

13.
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   

14.
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   

15.
The electrical conductivities of UO2+x. ThO2 and their solid solutions, in thermodynamic equilibrium with the gas phase, were measured as a function of temperature, and of oxygen partial pressure in the temperatnre range 800 to 1200°C. The slope of the plot log α versus 1/T for UO2+x and UO2-rich solid solutions exhibits a single region, whereas in the ThO2-rich solid solutions it exhibits two regions. The pressure dependence of the conductivity (σ) in the UO2-rich solid solutions can be represented by σ ∝ [Oi] ∝ po212 in the range of 0.01 < x < 0.1. Here, Oi is an interstitial oxygen and po2 the partial pressure of oxygen, and it varies with the ThO2 content. At greater deviation from stoichiometry (x ? 0.1) the presence of U4O9 or (Th U)4O9 phases influences the conductivity data. In ThO2 or ThO2-rich solid solutions. P-type conduction at high oxygen pressures is interpreted as arising from the incorporation of excess oxygen into oxygen vacancies.  相似文献   

16.
17.
The high-temperature specific heat of solid UO2, ThO2, and Al2O3 can be represented by an equation of the form Cp(s) = 3nRF(?D/T) + dT3, (1) where ?D is the Debye temperature, F(?D/T) is the Debye function, d represents contributions of the anharmonic vibrations within the lattice, and n denotes the number of atoms per molecule. In the liquid the corresponding equation is Cp(1) = 3nRF(?D/T) + hT2, (2) where h is the anharmonic term. It is shown that for Al2O3 and UO2, where experimental data for the liquid phase are also available, dh has the same value, Indicating that both materials behave identically. If we compare the thermodynamic relationship Cp ? Cv = Vα2KT, (3) where V is the volume, α the volume expansion coefficient, and K the bulk modulus, with equation (1), It follows that d must be equal to 2KT2; the value of 2KT2 is calculated in the temperature region where d was obtained; within experimental error they are equal.  相似文献   

18.
The formation of U2C3 by the reaction of UC2 with UO2 has been studied by chemical and X-ray analyses at temperatures between 1400 and 1700 °C in vacuo. The reaction is represented by 7 UC2 + UO2 → 4 U2C3 + 2 CO.  相似文献   

19.
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   

20.
The creep of UO2 containing small additions of Nb2O5 has been investigated in the stress range 0.5–90 MN/m2 at temperatures between 1422 and 1573 K. The functional dependence of the creep rate of five dopant concentrations up to 0.8 mol% Nb2O5 has been examined and it was established that in all the materials the secondary creep rate could be represented by the equation /.εkT = nexp(?Q/RT), where /.ε is the steady state creep rate per hour, Q the activation energy and A and n are constants for each material. It was observed that Nb2O5 additions can cause a dramatic increase in the steady state creep rate as long as the niobium ion is maintained in the Nb5+ valence state. Material containing 0.4 mol% Nb2O5 creeps three orders of magnitude faster than the pure material.Analysis of the results in terms of grain size compensated viscosity suggest that, like “pure” UO2, the creep rate of Nb2O5 doped fuel is diffusion-controlled and proportional to the reciprocal square of the grain size. A model is developed which suggests that the increase in creep rate results from suppression of the U5+ ion concentration by the addition of Mb5+ ions, which modifies the crystal defect structure and hence the uranium ion diffusion coefficient.  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