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.
Measurements of low-frequency internal friction and electron microscope observations were made on neutron-irradiated vanadium with various oxygen contents. Irradiation was carried out at about 60°C to a fast fluence of 2 × 1017 or 5 × 1019 n/cm2 (E ? 1 MeV). The oxygen Snoek damping was decreased by irradiation and post-irradiation annealing below 200 or 250° C, while it began to recover by annealing above this temperature. Complete recovery was attained by 30 min anneal at 450°C in the case of the lower fluence, whereas in the other case it was not observed after the same treatment. The results of electron microscope observations were consistent with those of internal friction measurements. The specimens irradiated to 5 × 1019 n/cm2 showed an abnormal peak after annealing above 250°C near the nitrogen Snoek temperature. The height of this peak, P?1max, was expressed as P?1max ∝ exp (2.72 × 103/RT) Q?1max, where Q?1max the heiβht of the oxygen Snoek damping after each annealing. The mechanism for radiation-anneal hardening and the abnormal peak were considered in the light of these experiments.  相似文献   

2.
Three massive samples of pyrocarbon were irradiated at 1100°C for a maximum fast-neutron dose of 1.6 × 1021 DNE. They were subjected to stresses in the range 1.33 × 102–2 × 102 Kg/cm2. The pyrocarbon was deposited from methane in a rotating furnace. Its density, its isotropy, its structure according to X-rays and TEM relate closely to its homologue deposited from methane in fluidised conditions. A study of creep under irradiation showed that a brief stage of primary creep is followed by a stage which is linear with respect to both stress and fast-neutron dose. Creep is thus well represented by an expression of the form ? = Kσφ, where K is 2 × 10?25 (Kg · cm?2. DNE)?1, which is a value ten times greater than previously estimated. Irradiation is accompanied by densification, a slight increase in anisotropy and a reduction in Lc (apparent crystallite size measured along the c axis). The variation of these parameters with dose does not, however, differ appreciably between the three creep samples and the unstressed sample.  相似文献   

3.
Previously published data on the final stage sintering kinetics of stoichiometric uranium dioxide are correlated with a reinterpretation of low-stress creep behaviour of identical material (data on both processes by the present authors). For both processes the rate-controlling diffusional flux is considered to be that of uranium ions along grain boundaries. The effective diffusion coefficient for uranium ion diffusion along grain boundaries, DGB, is estimated to be: DGB = 1.38(÷x5) × 10?6 exp ? [(2.39 ± 0.8) × 105/8.31T] m2/s. Comparisons are made between this value and those previously measured by radio-tracer methods.  相似文献   

4.
The sessile drop method was used for the determination of the density in liquid state. The results for stainless steel 1.4970 using uranium dioxide as substrate material in the temperature range 1690 K (liquidus temperature T1) < T < 2120 K are ρ = 6.82 × 103 ? 10.25 × 10?1 (T ? T1) kg/m3, and α = 1.50 × 10?4K?1. Below 1690 K the linear thermal expansion is given by Δl/l0 = 0.00204 + 7.110 × 10?6 T + 7.734 × 10?9 T2. Using the same method but not correlated with the density measurements the following interfacial properties of the system UO2-stainless steel have been determined: surface energy of liquid steel γLv = 1.19 ? 0.57 × 10?3 (T ? T1) J/m2 and interfacial energy of liquid steel against UO2γSL = 1.57 ? 2.01 × 10?3 (T ? T1) J/m2, the results yield a contact angle θ = 0° at T= 2515 K. Using literature data for the compressibility of liquid UO2, an estimate of the surface energy of UO2 in liquid state was performed. The estimated value at the melting point is: γLV = 0.522 J/m2. The mean value of the experimental data given by several authors is 0.513 ± 0.085 J/m2. The estimated temperature dependence of the surface energy of liquid UO2 is given by dγLV/dT = ?0.19 × 10?3J/m2.  相似文献   

5.
A study was performed of the diffusion in α-thorium of fission products representing impurity atoms with a diversity of size and valance differences with respect to the solvent lattice. The atoms were recoil injected into thorium disks. Diffusion coefficients were determined for 133Xe by monitoring its release during annealing, and for the other isotopes by post-annealing concentration profile analysis. The Arrhenius constants do(cm2/sec), resp. Q(kcal/mole) were obtained for the diffusion coefficients where. D = Doexp(?Q/RT);99Mo: 7.6 × 10?4 and 37.6; 132Te: 1.32 × 106 and 95.6; 133I: 2.7 × 10?1 and 66.2; 133Xe: 3.6 × 102 and 82.3; 140Ba: 2.3 × 10?2 and 59.4; 141Ce ? 143Ce: 1.8 × 10?2 and 60.0. The fission product diffusion behavior, in general, fit either the vacancy or the substitutional-interstitial diffusion mechanisms for impurity atoms in a fcc metal. Both valence and ionic radius correlations were found. The data indicate low rates of diffusion for the operating temperatures at which α-thorium-based fuel might be used.  相似文献   

6.
Diffusion of carbon in zirconium, zircaloy-2 and Zr- 2.5% Nb has been studied in the temperature range 873–1523K for zirconium and zircaloy-2 and 753–1523K for Zr-2.5% Nb alloy, using the residual activity technique. The diffusivities (in m2/s) in the α and β phases could be represented by DC/α-Zr(873–1123K) = (2.00 ± 0.37) × 10?7 exp [?(151.59 ± 2.51)RT]DC/α-Zircaloy-2 (873–1043K) = (1.41 ± 0.32) × 10?7 exp [?(158.99 ± 3.14)RT]DC/α-Zr-Nb-alloy (753–873K) = (4.68 ± 0.88) × 10?7 exp [?(159.98 ± 2.91)RT]DC/β Zr ((1143–1523K) = (8.90 ± 1.60) × 10?6 exp [?(133.05 ± 1.46)RT]DC/β Zircaloy-2 (1263–1523K) = (2.45 ± 0.61) × 10?5 exp [?(150.29 ± 1.72)RT]DC/β Zr-Nb alloy (1143–1523) = (1.70 ± 0.42) × 10?5 exp [?(158.20 ± 2.09)RT]The activation energies are given in kJ/mole. In the phase transition region, the diffusivities could be represented by the empirical relation: D = Dα · Dβ, where Cα, Cβ are the concentrations of the two phases in the alloy and Dα, Dβ are the extrapolated values of diffusion co-efficients in the α and β phases respectively.The results have been explained in terms of the interstitial mechanism of diffusion.  相似文献   

7.
Steady-state creep rates of as-received zircaloy-4 fuel cladding have been determined from 940 to 1073 K in the α-Zr range, from 1140 to 1190 K in the mixed (α + β) phase region and from 1273 to 1873 K in the β-Zr phase region. Strain rates of between 10?6 and 10?2/s were determined under constant uniaxial load conditions. Assuming that creep rates can be described by a power law-Arrhenius equation, the creep rate for α-phase zircaloy-4 is given by: gess? = 2000 σ5.32exp(?284 600/kT) s?1; for the β-phase zircaloy-4 by: gess?= 8.1 σ3.79exp(?142 300/kT) s?1; and for the mixed (α + β) phase of zircaloy-4 (for creep rates ?3 × 10?3 s?1) by: gess?= 6.8 × 10?3 σ1.8exp(?56 600/kT) s?1. For the both the α and β phases, the activation energies for creep are in agreement with those of self-diffusion. For the mixed (α + β) phase region, the low creep rate range is controlled by grain boundary sliding at the α/(α + β) phase boundary.  相似文献   

8.
Tensile tests were carried out on Zircaloy-4 over the temperature range 298–798 K. Yield stress values at the strain rates 1.33 × 10?4s?1 and 6.67 × 10?4s?1 were used to determine the activation parameters. A peak in activation volume (Vapp = 3100 b3) was observed at about 690 K; outside this temperature range the activation volumes became almost independent of temperature (Vapp = 200?300 b3). The peak in activation volume was explained in terms of a basic rate controlling mechanism and dynamic strain aging. This analysis indicated that the peak could be ascribed to the negative value of the strain rate sensitive solute strengthening term M and that the mechanism based on the non-conservative motion of jogs appeared to be more favored as the basic rate controlling mechanism of Zircaloy-4 than an impurity mechanism  相似文献   

9.
The oxidation kinetics in air of Van Arkel hafnium were studied between 750 and 950°C. Comparison of the results obtained by thermogravimetry and microhardness measurements has allowed us to characterise the overall oxidation kinetics which are of parabolic type and also the two separate kinetics, likewise parabolic, of growth of the oxide film and of dissolution of oxygen in the metal underlying the oxide. For each of these three kinetics the rate constants are respectively given by the following relations: Kp = 3.1 × 108 exp (?46 000/RT), K1 = 1.4 × 106 exp (?36 000/RT), K2 = 3.4 × 1010 exp (?59 000/RT).These results are in agreement with a classical diffusion model of oxygen, both in the oxide and in the metal, set up on the basis of Wagner's general theory. The activation energy of diffusion of oxygen in hafnium is 53.0 ± 2 kcal/mole.  相似文献   

10.
The nuclear magnetic resonance (NMR) from ion-implanted 3He atoms has been observed in 3He+-bombarded palladium. Two 1 μm palladium thin films, one on each side of a copper foil substrate, were bombarded at 75°C with 75 and 140 keV 3He+-ions to a fluence of ~10183He+/cm2 for each energy. Spin-lattice (T1) and spin-spin (T2) relaxation times were measured at 10, 20, and 35 MHz in the temperature range 1–4K. The 3He nuclear relaxation data indicate that the implanted atoms after the 75°C bombardment are situated in small clusters within the Pd thin films.  相似文献   

11.
12.
The diffusion coefficients of 7Be in both α and β phases of Zr, are reported. The temperature dependence of the diffusion coefficient in the α phase may be expressed by DBeα-Zr = 0.33 exp(?31900/RT) cm2/s. The measured values in the β phase are in agreement with previously literature reported data, which give a temperature dependence expressed by DBeβ-Zr = 8.33 × 10?2 exp(?31800/RT) cm2/s. Be diffusion in Zr, which is consistent with an interstitial-like behavior, is analyzed in terms of the Anthony and Turnbull conditions, and atomic size criteria. It is concluded that the latter is a very important parameter when assessing the possibility of significant interstitial-like dissolution.  相似文献   

13.
300 keV Ar+ ions are bombarded on the surface of Mo single crystals to doses of 1 × 1017, 6.2 × 1017, 1 × 1018 and 2.8 × 1018 ions/cm2. After bombardment with 6.2 × 1017 ions/cm2, blistering is observed after room-temperature aging of less than 100 days, in spite of the large sputtering yield. Disappearance of the formed blisters and concurrent surface roughening are observed with further aging. For higher dose bombardments, only surface roughening is observed without prior formation of blisters. Distribution of the injected ions associated with large sputtering yield is derived. From this distribution, the critical amount of injected ions required for blistering to occur is estimated to be equal to or less than 3.4 × 1017 ions/cm2.  相似文献   

14.
The nonstoichiometric composition of Cr2O3±x was measured by means of thermogravimetry in the range of 1173 ≦ T/K ≦ 1318 and 10?15 ≦ PO2/Pa ≦ 105. The compositional deviation from stoichiometry, x, in the hyperstoichiometric Cr2O3+x phase was observed to be smaller than 2 × 10?4, irrespective of temperatures, provided that the hyperstoichiometric Cr2O3+x exists. The existence of the hypostoichiometric Cr2O3?x phase was first established in this study in the region of low oxygen partial pressure below 10?5 Pa. From the oxygen partial pressure dependence of x in Cr2O3?x, the defect structure was discussed with the neutral chromium interstitials in the composition near stoichiometry and with the triply charged ones far from stoichiometry. The partial molar enthalpy and entropy of oxygen of Cr2O3?x showed the complex compositional dependences, suggesting the change of the type of the predominant defect.  相似文献   

15.
The solute diffusion at infinite dilution of 198Au and 110mAg in cubic phases of Pu has been studied using the serial sectroning method. The solute diffusion coefficients in the b.c.c. ? phase can be expressed by: DAu?Pu = 5,7 × 10?5 exp(?10300/RT) cm2/s and DAg?Pu = 4,9 × 10?5 exp(?9600/RT) cm2/s. The solute diffusion mechanism is interstitial of the dissociative type in both cases. These experiments confirm the activated interstitial model which has been proposed for self diffusion of ?Pu. Indeed the solute diffusion coefficients of Au and Ag are near of the self diffusion coefficients of Pu. The mechanisms are therefore interstitial in both cases. In the f.c.c. δ phase of Pu where self diffusion takes place by a vacancy mechanism, the solute diffusion coefficients of Au and Ag are near of the self diffusion coefficients of δ Pu. Solute diffusion takes place also by a vacancy mechanism. On the other hand, the extrapolation at infinite dilution of experiments of solute diffusion of Cu in ?Pu (Matano-Wagner coupling) gives the following results: DCu?Pu = 1 × 10?3 exp(?12300/RT) cm2/s. The solute diffusion mechanism is interstitial of the dissociative type. In the ? phase the smaller the atomic radius the faster the migration: rCo < rCu < r?Pu < rAg = rAu, and DCo?Pu > DCu?Pu >DPu?PU > DAg?Pu ≈ DAu?Pu.  相似文献   

16.
A single crystal of crystal bar Zr was irradiated, unstressed, at 570 K in a fast (> 1 MeV) neutron flux of 5.5 × 1016n/m2-s. After a dose of 6 × 1023n/m2 a tensile stress of 25 MPa was applied during a period of steady reactor power. The loading strain was an order of magnitude smaller than that observed when an identical, unirradiated, crystal was loaded to the same stress. There followed a period of primary creep during which the creep rate decreased to a value of 5 × 10?6h?1 in the first 24 hours of the test. For the final 2000 hours of the test the specimen was observed to creep at a rate of 1 × 10?6h?1 when the reactor was at full power. During shutdowns, the creep rate decreased with time. The results will be discussed and compared with predictions from current theories for the mechanism of irradiation enhanced creep in light of the micro-structures observed.  相似文献   

17.
The as-irradiated microstructure of molybdenum, irradiated in the EBR-II reactor at six different temperatures in the range 430–1000°C (0.24–0.44 Tm) to a fast neutron fluence of ≈ 1 × 1022 n · cm?2 (E > 1 MeV), has been characterized as black spot clusters, loops, rafts, voids (random and ordered) and dislocations. Present results show that both the void density, Nv and the void size, dv, are independent of irradiation temperature in the range 430–700° C. Above the 700° C irradiation temperature the void density decreases and the void size increases exponentially with increasing irradiation temperature and they have been expressed empirically as Nv = 3.6 × 1020exp (?26.9 T/Tm), dv = 1.5 exp (9.44 T/Tm), where T/Tm is the irradiation temperature presented as a fraction of the melting point. The void density of all available published data has been used to show that the void density is (a) a strong function of irradiation temperature for a constant number of displacements per atom (dpa) and (b) a function of reactor power and spectrum when normalized to dpa.  相似文献   

18.
Vanadium samples were neutron irradiated at the reactor ambient temperature to fluences in the range from 2.0 × 107 to 1.0 × 1020 n/cm2 (En ? 1 MeV). The radiation hardening measured at the ambient temperature increased linearly with the square root of the neutron fluence, up to a fluence of about 2.5 × 1019 n/cm2, to approximately 25 kg/mm2 for the highest fluence. The radiation-anneal hardening phenomenon was clearly observed in samples irradiated at a low fluence (2.0 × 1017 and 1.0 × 1018 n/cm2) and the hardening was accompanied by changes in the density and size distribution of the radiation-produced defect clusters. The radiation hardening induced during irradiation to 1.0 × 1020 n/cm2 recovered monotonically as the annealing temperature increased. Defect clusters invisible in the electron microscope played an important role in the radiation and anneal hardening except when radiation hardening was induced at the highest fluence.  相似文献   

19.
Temperature profiles similar to those existing in fuel rods under irradiation have been simulated by passing electrical current through cylindrical pellets. The comparison between calculated and measured temperatures in pellets heated in a thermal gradient shows: (1) The values of thermal conductivity obtained by different authors in isothermal experiments and extrapolated to temperatures up to 2700°C are not in agreement. Therefore, the calculation of the temperature of the fuel leads to errors which vary between 1 and 16% depending on the data for λ used. For central temperatures above 1900°C the values of Schmidt better suit the calculations, especially if the oxygen content of the fuel is smaller than 2.00. (2) The published data of electrical conductivity are in open disagreement. The values of activation energy are generally higher than those deduced from the present investigation. It has been assumed that the activation energy E in the equation σ = A exp(?E(T)/kT) varies with temperature as E(T)= E0(1?1.94 × 10?4T) when E0= 0.58 eV if O/M = 2.00, and E(T) = E0(1?2.21× 10?4T) when E0 =0.91 eV if O/M = 1.94.  相似文献   

20.
The post-irradiation annealing behavior of β-SiC for use as a monitor of irradiation temperature is discussed. Powder and rods of polycrystalline β-SiC were irradiated to 1.5 × 1017 to 5.0 × 1019 n/cm2 (E > 0.18 MeV) at temperatures between 290 and 500°C. The estimated temperatures deduced from the changes in lattice constant and specific electric resistivity during progressive annealing, and from thermal expansion measurement by high-temperature X-ray diffraction agreed with values determined by means of a thermocouple. Thermal expansion measurement in a conventional dilatometer resulted in an over estimate of the irradiation temperature, and further improvement of this method is required for experimental application.  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