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 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   

2.
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   

3.
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   

4.
Density functional theory (DFT) calculations of fission product (Xe, Sr, and Cs) incorporation and segregation in alkaline earth metal oxides, HfO2 and UO2 oxides, and the MgO/(U, Hf, Ce)O2 interfaces have been carried out. These calculations demonstrate that the fission product incorporation energies in MgO are higher than in HfO2. However, this trend is reversed or reduced for alkaline earth oxides with larger cation sizes. In the case of UO2, the calculations were performed using spin polarization and with a Hubbard U term characterizing the on-site Coulomb repulsion between the localized 5f electrons. The fission product solution energies in bulk UO2 ± x have been calculated as a function of non-stoichiometry x, and were compared to that in MgO. The solution energies of fission products in MgO are substantially higher than in UO2 ± x, except for the case of Sr in hypostoichiometric UO2. Due to size effects, the thermodynamic driving force of segregation for Xe and Cs from bulk MgO to the MgO/fluorite interface is strong. However, this driving force is comparatively weak for Sr.  相似文献   

5.
(U, Pu) mixed oxides, (U1−yPuy)O2−x, with y = 0.21 and 0.28 are being considered as fuels for the Prototype Fast Breeder Reactor (PFBR) in India. The use of urania-plutonia solid solutions in PFBR calls for accurate measurement of physicochemical properties of these materials. Hence, in the present study, oxygen potentials of (U1−yPuy)O2−x, with y = 0.21 and 0.28 were measured over the temperature range 1073-1473 K covering an oxygen potential range of −550 to −300 kJ mol−1 (O/M ratio from 1.96 to 2.000) by employing a H2/H2O gas equilibration technique followed by solid electrolyte EMFmeasurement. (U1−yPuy)O2−x, with y = 0.40 is being used in the Fast Breeder Test Reactor (FBTR) in India to test the behaviour of fuels with high plutonium content. However, data on the oxygen potential as well as thermal conductivity of the mixed oxides with high plutonium content are scanty. Hence, the thermal diffusivity of (U1−yPuy)O2, with y = 0.21, 0.28 and 0.40 was measured and the results of the measurements are reported.  相似文献   

6.
In order to simulate the effects of burnable poison doping on the fission fragment damage of UO2 nuclear fuels, Er2O3-doped CeO2 pellets were irradiated with 200 MeV Xe14+ ions. The irradiation effect was measured by means of X-ray diffraction (XRD). The expansion of lattice and the disordering of atomic arrangement due to the irradiation become more remarkable with increasing the concentration of the Er2O3 dopant.  相似文献   

7.
Inert matrix fuels are an important component of advanced nuclear fuel cycles, as they provide a means of utilizing plutonium and reducing the inventory of ‘minor’ actinides. We examine the neutronic and thermal characteristics of MgO-pyrochlore (A2B2O7: La2Zr2O7, Nd2Zr2O7 and Y2Sn2O7) composites as inert matrix fuels in boiling water reactors. By incorporating plutonium with resonance nuclides, such as Am, Np and Er, in the A-site of pyrochlore, the kinfvs. burn-up curves are shown to be similar to those of UO2, although the Doppler coefficients are less negative than UO2. The Pu depletion rates are 88-90% (239Pu) and 54-58% (total Pu) when the inert matrix fuels experience a burn-up equivalent of 45 GWd/tHM UO2. Because of the high thermal conductivity of MgO, the center-line temperatures of the MgO-pyrochlore composites at 44.0 kW/m are lower than those of UO2 pellets. After burn-up, the A-site cation composition is 15-35 at.% lower than that of the B-site cations in pyrochlore (e.g., A1.84B2.17O7.00) due to the fission of Pu in the A-site and the presence of fission product elements in the A- and B-sites of the pyrochlore structure.  相似文献   

8.
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   

9.
Particles of UO2+x (x≅0.16 ± 0.06) exposed to the atmosphere react by oxidation and formation of complexes (hydrates, hydroxides and carbonates). Surface reactions alter and erode the UO2 particles. This paper outlines results for measurements of oxidation rates on uranium oxide particles using in situ photoluminescence spectroscopy (PL), X-ray photoelectron spectroscopy (XPS) and secondary ion mass spectrometry (SIMS). Phosphorescence spectra observed during oxidation of UO2+x were attributed to U(VI) in uranyl-type coordination and in octahedral coordination. Uranyl-type spectra formed during wet oxidation of UO2+x, and U(VI) octahedral spectra formed during dry oxidation of UO2+x. The uranyl-type species, although more stable, is more kinetically labile for vacuum reduction than is the octahedral U(VI). Oxidation of U(IV) species are diffusion controlled. Vacuum reduction of uranyl U(VI) in UO3 follows a field-enhanced cationic diffusion rate law, while re-oxidation follows a diffusion rate law. Post-oxidation core and valence band XPS and SIMS measurements provided qualitative and quantitative measures of uranium oxidation states near uranium oxide surfaces.  相似文献   

10.
UO2 and (U, Pu)O2 solid solutions (the so-called MOX) nowadays are used as commercial nuclear fuels in many countries. One of the safety issues during the storage of these fuels is related to their self-irradiation that produces and accumulates point defects and helium therein.We present density functional theory (DFT) calculations for UO2, PuO2 and MOX containing He atoms in octahedral interstitial positions. In particular, we calculated basic MOX properties and He incorporation energies as functions of Pu concentration within the spin-polarized, generalized gradient approximation (GGA) DFT calculations. We also included the on-site electron correlation corrections using the Hubbard model (in the framework of the so-called DFT + U approach). We found that PuO2 remains semiconducting with He in the octahedral position while UO2 requires a specific lattice distortion. Both materials reveal a positive energy for He incorporation, which, therefore, is an exothermic process. The He incorporation energy increases with the Pu concentration in the MOX fuel.  相似文献   

11.
The melting behavior of MgO-based inert matrix fuels containing (Pu,Am)O2−x ((Pu,Am)O2−x-MgO fuels) was experimentally investigated. Heat-treatment tests were carried out at 2173 K, 2373 K and 2573 K each. The fuel melted at about 2573 K in the eutectic reaction of the Pu-Am-Mg-O system. The (Pu,Am)O2−x grains, MgO grains and pores grew with increasing temperature. In addition, Am-rich oxide phases were formed in the (Pu,Am)O2−x phase by heat-treatment at high temperatures. The melting behavior was compared with behaviors of PuO2−x-MgO and AmO2−x-MgO fuels.  相似文献   

12.
In a deep repository for spent nuclear fuel, U(VI)(aq) released upon dissolution of the fuel matrix could, in reducing parts of the system, be converted to U(IV) species which might coalesce and form nanometer-sized UO2 particles. This type of particles is expected to have different properties compared to bulk UO2(s). Hence, their properties, in particular the capacity for oxidant consumption, must be investigated in order to assess the effects of formation of such particles in a deep repository. In this work, methods for radiation chemical synthesis of nanometer-sized UO2 particles, by electron- and γ-irradiation of U(VI) solutions, are presented. Electron-irradiation proved to be the most efficient method, showing high conversions of U(VI) and yielding small particles with a narrow size distribution (22-35 nm). Stable colloidal suspensions were obtained at low pH and ionic strength (pH 3, I = 0.03). Furthermore, the reactivity of the produced UO2 particles towards H2O2 is investigated. The U(IV) fraction in the produced particles was found to be ∼20% of the total uranium content, and the results show that the UO2 nanoparticles are significantly more reactive than micrometer-sized UO2 when it comes to H2O2 consumption, the major part of the H2O2 being catalytically decomposed on the particle surface.  相似文献   

13.
The thermal conductivities of (U0.68Pu0.30Am0.02)O2.00−x solid solutions (x = 0.00-0.08) were studied at temperatures from 900 to 1773 K. The thermal conductivities were obtained from the thermal diffusivities measured by the laser flash method. The thermal conductivities obtained experimentally up to about 1400 K could be expressed by a classical phonon transport model, λ = (A + BT)−1, A(x) = 3.31 × x + 9.92 × 10−3 (mK/W) and B(x) = (−6.68 × x + 2.46) × 10−4 (m/W). The experimental A values showed a good agreement with theoretical predictions, but the experimental B values showed not so good agreement with the theoretical ones in the low O/M ratio region. From the comparison of A and B values obtained in this study with the ones of (U,Pu)O2−x obtained by Duriez et al. [C. Duriez, J.P. Alessandri, T. Gervais, Y. Philipponneau, J. Nucl. Mater. 277 (2000) 143], the addition of Am into (U, Pu)O2−x gave no significant effect on the O/M dependency of A and B values.  相似文献   

14.
Solid state reactions of UO2, ThO2, PuO2 and their mixed oxides (U, Th)O2 and (U, Pu)O2 were carried out with sodium nitrate upto 900 °C, to study the formation of various phases at different temperatures, which are amenable for easy dissolution and separation of the actinide elements in dilute acid. Products formed by reacting unsintered as well as sintered UO2 with NaNO3 above 500 °C were readily soluble in 2 M HNO3, whereas ThO2 and PuO2 did not react with NaNO3 to form any soluble products. Thus reactions of mixed oxides (U, Th)O2 and (U, Pu)O2 with NaNO3 were carried out to study the quantitative separation of U from (U, Th)O2 and (U, Pu)O2. X-ray diffraction, X-ray fluorescence, thermal analysis and chemical analysis techniques were used for the characterization of the products formed during the reactions.  相似文献   

15.
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   

16.
ThxU1−xO2+y binary compositions occur in nature, uranothorianite, and as a mixed oxide nuclear fuel. As a nuclear fuel, important properties, such as the melting point, thermal conductivity, and the thermal expansion coefficient change as a function of composition. Additionally, for direct disposal of ThxU1−xO2, the chemical durability changes as a function of composition, with the dissolution rate decreasing with increasing thoria content. UO2 and ThO2 have the same isometric structure, and the ionic radii of 8-fold coordinated U4+ and Th4+ are similar (1.14 nm and 1.19 nm, respectively). Thus, this binary is expected to form a complete solid solution. However, atomic-scale measurements or simulations of cation ordering and the associated thermodynamic properties of the ThxU1−xO2 system have yet to be determined. A combination of density-functional theory, Monte-Carlo methods, and thermodynamic integration are used to calculate thermodynamic properties of the ThxU1−xO2 binary (ΔHmix, ΔGmix, ΔSmix, phase diagram). The Gibbs free energy of mixing (ΔGmix) shows a miscibility gap at equilibration temperatures below 1000 K (e.g., Eexsoln = 0.13 kJ/(mol cations) at 750 K). Such a miscibility gap may indicate possible exsolution (i.e., phase separation upon cooling). A unique approach to evaluate the likelihood and kinetics of forming interfaces between U-rich and Th-rich has been chosen that compares the energy gain of forming separate phases with estimated energy losses of forming necessary interfaces. The result of such an approach is that the thermodynamic gain of phase separation does not overcome the increase in interface energy between exsolution lamellae for thin exsolution lamellae (10 Å). Lamella formation becomes energetically favorable with a reduction of the interface area and, thus, an increase in lamella thickness to >45 Å. However, this increase in lamellae thickness may be diffusion limited. Monte-Carlo simulations converge to an exsolved structure [lamellae || ] only for very low equilibration temperatures (below room temperature). In addition to the weak tendency to exsolve, there is an ordered arrangement of Th and U in the solid solution [alternating U and Th layers || {1 0 0}] that is energetically favored for the homogeneously mixed 50% Th configurations. Still, this tendency to order is so weak that ordering is seldom reached due to kinetic hindrances. The configurational entropy of mixing (ΔSmix) is approximately equal to the point entropy at all temperatures, indicating that the system is not ordered.  相似文献   

17.
The thermal conductivities of δ′-, δ-, δ+ε-, and ε-phase hafnium hydrides and deuterides with various hydrogen isotope concentrations (HfHx, 1.48 ? x ? 2.03; HfDx, 1.55 ? x ? 1.94) were evaluated within the temperature range of 290-570 K from the measured thermal diffusivity, calculated specific heat, and density. The thermal conductivities of δ′-, δ-, δ+ε-, and ε-phase HfHx and HfDx are independent of the temperature within the range 300-550 K and are in the range 0.15-0.22 W/cm K and 0.17-0.23 W/cm K, respectively; these values are similar to and lower than the observed thermal conductivities of α-phase Hf. The experimental results for the electrical resistivities of δ′-, δ-, δ+ε-, and ε-phase HfHx and HfDx and the Lorenz number corresponding to the electronic conduction, obtained from the Wiedemann-Franz rule, indicated that heat conduction due to electron migration significantly influences the thermal conductivity values at high temperatures. On the other hand, heat conduction due to phonon migration significantly affects the isotope effects on the thermal transport properties.  相似文献   

18.
The thermal conductivity of nuclear fuels such as UO2+x and (U,Pu)O2−x has been calculated by the molecular dynamics (MD) simulation in terms of oxygen stoichiometric parameter x, temperature and Pu content. In the present study, the MD calculations were carried out in both equilibrium (EMD) and nonequilibrium (NEMD) systems. In the EMD simulation, the thermal conductivity was defined as the time-integral of the correlation function of heat fluxes according to the Green-Kubo relationship. Meanwhile, in the homogeneous NEMD, it was given by the ratio of the time-averaged heat flux to the perturbed external force subjected to each particle in the simulated cell. NEMD, as compared with EMD, gave somewhat precise results efficiently. Furthermore, both MD calculations showed that the thermal conductivity of these oxide fuels decreased with increase of temperature and defects, i.e. excess oxygen or vacancy, and was rather insensitive to Pu content for the stoichiometric fuel.  相似文献   

19.
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   

20.
Results of the investigation of the FeO1.5-UO2+x-ZrO2 system in air are presented. The eutectic position and the content of the phases crystallized at this point have been determined. The temperature and the composition of the ternary eutectic are 1323 ± 7 °С and 67.4 ± 1.0 FeO1.5, 30.5 ± 1.0 UO2+x, 2.1 ± 0.2 ZrO2 mol.%, respectively. The solubilities of FeO1.5 and ZrO2 in the UO2+x(FeO1.5, ZrO2) solid solution correspond to respectively 3.2 and 1.1 mol.%. The solubilities of UO2 and ZrO2 in FeO1.5 are not significant. The existence of a solid solution on the basis of U(Zr)FeO4 compound is found. The ZrO2 solubility in this solid solution is 7.0 mol.%.  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