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.
Thermal expansions of UO2 and a simulated fuel with fission products forming a solid solution were studied using a dilatometer in the temperature range from 298 to 1800 K. The densities of the UO2 and the simulated fuel used in the measurements were 10.43 g · cm−3 (95.2% of theoretical density (TD)) and 10.35 g · cm−3 (95.6% of TD), respectively. The linear thermal expansion of the simulated fuel is higher than that of UO2, and the difference between this fuel and UO2 increases monotonically with temperature. The average linear thermal expansion coefficients of UO2 and the simulated fuel are 1.09× 10−5 and 1.23×10−5 K−1, respectively. As the temperature increases to 1800 K, the relative densities of UO2 and the simulated fuel decrease to 95.1 and 94.7% of their initial densities at 298 K.Paper presented at the Seventh Asian Thermophysical Properties Conference, August 23–28, 2004, Hefei and Huangshan, Anhui, P. R. China.  相似文献   

2.
As a part of the DUPIC (direct use of spent PWR fuel in CANDU reactors) fuel development program, the thermal expansion of simulated spent fuel pellets with dissolved fission products has been studied by using a thermo-mechanical analyzer (TMA) in the temperature range from 298 K to 1773 K to investigate the effects of fission products forming solid solutions in a UO2 matrix on the thermal expansions. Simulated fuels with an equivalent burn-up of (30 to 120) GWd/tU were used in this study. The linear thermal expansions of the simulated fuel pellets were higher than that of UO2, and the difference between these fuel pellets and UO2 increased monotonically with temperature. For the temperature range from 298 K to 1773 K, the values of the average linear thermal expansion coefficients for UO2 and simulated fuels with an equivalent burn-up of (30, 60, and 120) GWd/tU are 1.19 × 10−5 K−1, 1.22 × 10−5 K−1, 1.26 × 10−5 K−1, and 1.32 × 10−5 K−1, respectively.  相似文献   

3.
The thermal diffusivity of a simulated fuel with fission products forming a solid solution was measured using the laser-flash method in the temperature range from room temperature to 1673 K. The density and the grain size of the simulated fuel with the solid solutions used in the measurement were 10.49 g · cm−3 (96.9% of theoretical density) at room temperature and 9.5 μm, respectively. The diameter and thickness of the specimens were 10 and 1 mm, respectively. The thermal diffusivity decreased from 2.108 m2 · s−1 at room temperature to 0.626 m2 · s−1 at 1673 K. The thermal conductivity was calculated by combining the thermal diffusivity with the specific heat and density. The thermal conductivity of the simulated fuel with the dissolved fission products decreased from 4.973 W · m−1 · K−1 at 300 K to 2.02 W · m−1 · K−1 at 1673 K. The thermal conductivity of the simulated fuel was lower than that of UO2 by 34.36% at 300 K and by 15.05% at 1673 K. The difference in the thermal conductivity between the simulated fuel and UO2 was large at room temperature, and decreased with an increase in temperature. Paper presented at the Seventeenth European Conference on Thermophysical Properties, September 5–8, 2005, Bratislava, Slovak Republic.  相似文献   

4.
The thermal conductivities of tin and lead in solid and liquid states have been determined using a nonstationary hot wire method. Measurements on tin and lead were carried out over temperature ranges of 293 to 1473 K and 293 to 1373 K, respectively. The thermal conductivity of solid tin is 63.9±1.3 Wm–1K–1 at 293 K and decreases with an increase in temperature, with a value of 56.6±0.9 Wm–1K–1 at 473 K. For solid lead, the thermal conductivity is 36.1±0.6 Wm–1K–1 at 293 K, decreases with an increase in temperature, and has a value of 29.1±1.1 Wm–1K–1 at 573 K. The temperature dependences for solid tin and lead are in good agreement with those estimated from the Wiedemann–Franz law using electrical conductivity values. The thermal conductivities of liquid tin displayed a value of 25.7±1.0 Wm–1K–1 at 573 K, and then increased, showing a maximum value of about 30.1 Wm–1K–1 at 673 K. Subsequently, the thermal conductivities gradually decreased with increasing temperature and the thermal conductivity was 10.1±1.0 Wm–1K–1 at 1473 K. In the case of liquid lead, the same tendency, as was the case of tin, was observed. The thermal conductivities of liquid lead displayed a value of 15.4±1.2 Wm–1K–1 at 673 K, with a maximum value of about 15.6 Wm–1K–1 at 773 K and a minimum value of about 11.4±0.6 Wm–1K–1 at 1373 K. The temperature dependence of thermal conductivity values in both liquids is discussed from the viewpoint of the Wiedemann–Franz law. The thermal conductivities for Group 14 elements at each temperature were compared.  相似文献   

5.
The density, the isobaric heat capacity, the surface tension, and the viscosity of liquid rhodium were measured over wide temperature ranges, including the supercooled phase, using an electrostatic levitation furnace. Over the 1820 to 2250 K temperature span, the density can be expressed as (T)=10.82×103–0.76(TT m ) (kgm–3) with T m =2236 K, yielding a volume expansion coefficient (T)=7.0×10–5 (K–1). The isobaric heat capacity can be estimated as C P (T)=32.2+1.4×10–3(TT m ) (Jmol–1K–1) if the hemispherical total emissivity of the liquid remains constant at 0.18 over the 1820 to 2250 K interval. The enthalpy and entropy of fusion have also been measured, respectively, as 23.0 kJmol–1 and 10.3 Jmol–1K–1. In addition, the surface tension can be expressed as (T)=1.94×103–0.30(TT m ) (mNm–1) and the viscosity as (T)=0.09 exp[6.4×104(RT)] (mPas) over the 1860 to 2380 K temperature range.  相似文献   

6.
Compounds Pb(BVUO6)2 nH2O (BV = P, As, V) were synthesized and studied by X-ray diffraction, IR spectroscopy, thermal analysis, and reaction calorimetry. The standard enthalpies and Gibbs energies of formation of the compounds at T = 298 K were determined, and processes involving these compounds were quantitatively characterized.Translated from Radiokhimiya, Vol. 46, No. 5, 2004, pp. 412–417.Original Russian Text Copyright © 2004 by Suleimanov, Chernorukov, Golubev.  相似文献   

7.
Several thermophysical properties of hafnium-3 mass % zirconium, namely the density, the thermal expansion coefficient, the constant pressure heat capacity, the hemispherical total emissivity, the surface tension and the viscosity are reported. These properties were measured over wide temperature ranges, including overheated and undercooled states, using an electrostatic levitation furnace developed by the National Space Development Agency of Japan. Over the 2220 to 2875 K temperature span, the density of the liquid can be expressed as L (T)=1.20×104–0.44(TT m ) (kgm–3) with T m =2504 K, yielding a volume expansion coefficient L (T)=3.7×10–5 (K–1). Similarly, over the 1950 to 2500 K span, the density of the high temperature and undercooled solid -phase can be fitted as S (T)=1.22×104–0.41(TT m ), giving a volume expansion coefficient S (T)=3.4×10–5. The constant pressure heat capacity of the liquid phase can be estimated as C PL (T)=33.47+7.92×10–4(TT m ) (Jmol–1K–1) if the hemispherical total emissivity of the liquid phase remains constant at 0.25 over the 2250 K to 2650 K temperature interval. Over the 1850 to 2500 K temperature span, the hemispherical total emissivity of the solid -phase can be represented as TS (T)=0.32+4.79×10–5(TT m ). The latent heat of fusion has also been measured as 15.1 kJmol–1. In addition, the surface tension can be expressed as (T)=1.614×103–0.100(TT m ) (mNm–1) and the viscosity as h(T)=0.495 exp [48.65×103/(RT)] (mPas) over the 2220 to 2675 K temperature range.  相似文献   

8.
Density Determination of Liquid Copper, Nickel, and Their Alloys   总被引:1,自引:0,他引:1  
A method for the determination of the density of electromagnetically levitated metallic liquids has been developed. This method employs an enlarged beam of parallel laser light to produce a shadow image of the sample. The shadow is recorded by a digital CCD-camera, and the images are analyzed using an edge detection algorithm. The circumference is fitted by Legendre polynomials that can be used for calculations of the volume of the sample. The method has been tested successfully on various alloys of copper-nickel (Ni x Cu y ), as well as on the pure elements, Cu and Ni. Densities were measured for each sample at different temperatures below and above the melting point, and a linear behavior was observed. At the melting point the densities for copper and nickel were 7.9 and 7.93gcm–3, respectively. For T=1270°C liquid copper has a density of 7.75gcm–3 which strongly increases up to roughly 8.1gcm–3 if a small amount (10–40 at.%) of nickel is added to the system. For nickel concentrations larger than 50at.% the density remains nearly constant.  相似文献   

9.
Solubility of UO2(NO3)2, TBP, and its degradation products in the organic and aqueous phases of the system TPB-H2O-HNO3-UO2(NO3)2 at 25–128°C over the uranyl nitrate concentration range 200–1200 g 1–1 is studied. The solubility of TBP and its degradation products in uranyl nitrate hexahydrate melt decreases in the order H2MBP > HDBP > TBP > H3PO4. At the boiling point of the melt, their solubility ranges from 100 to 0.5 g l–1. The solubility of uranyl nitrate in the TBP phase under the same conditions is caused by formation of complexes with a composition close to that of the monosolvate UO2(NO3) 2 TBP.Translated from Radiokhimiya, Vol. 46, No. 5, 2004, pp. 436–439.Original Russian Text Copyright © 2004 by Usachev, Markov.  相似文献   

10.
Densities of 1,1,1,2,3,3,3-heptafluoropropane (R227ea) have been measured with a computer-controlled high-temperature high-pressure vibrating-tube densimeter system (DMA-HDT) in the sub- and supercritical states. The densities were measured at temperatures from 278 to 473 K and pressures up to 30 MPa (overall 257 data points), whereby a density range between 285 and 1588 kgm–3 was covered. The uncertainty in the density measurement was estimated to be better than ±0.2 kgm–3. The experimental data of R227ea were correlated with a virial-type equation of state (EoS) and compared with published data. A comparison is also made with a recent wide-range dedicated equation of state for R227ea.  相似文献   

11.
The isochoric heat capacity C V of an equimolar H2O+D2O mixture was measured in the temperature range from 391 to 655 K, at near-critical liquid and vapor densities between 274.05 and 385.36 kgm–3. A high-temperature, high-pressure, nearly constant-volume adiabatic calorimeter was used. The measurements were performed in the one- and two-phase regions including the coexistence curve. The uncertainty of the heat-capacity measurement is estimated to be ±2%. The liquid and vapor one- and two-phase isochoric heat capacities, temperatures, and densities at saturation were extracted from the experimental data for each measured isochore. The critical temperature and the critical density for the equimolar H2O+D2O mixture were obtained from isochoric heat capacity measurements using the method of quasi-static thermograms. The measurements were compared with a crossover equation of state for H2O+D2O mixtures. The near-critical isochoric heat capacity behavior for the 0.5 H2O+0.5 D2O mixture was studied using the principle of isomorphism of critical phenomena. The experimental isochoric heat capacity data for the 0.5 H2O+0.5 D2O mixture exhibit a weak singularity, like that of both pure components. The reliability of the experimental method was confirmed with measurements on pure light water, for which the isochoric heat capacity was measured on the critical isochore (321.96 kgm–3) in both the one- and two-phase regions. The result for the phase-transition temperature (the critical temperature, T C, this work=647.104±0.003 K) agreed, within experimental uncertainty, with the critical temperature (T C, IAPWS=647.096 K) adopted by IAPWS.  相似文献   

12.
The behavior of 137Cs 131I radioactive aerosols at sorption in a column packed with Bekipor WB metal felt from argon, air, and steam-gas flow was studied to assess the possibility of formation of CsI CsOH and CsI3 in the course of CsI oxidative hydrolysis. The layer distribution of radionuclides in the column was analyzed as influenced by temperature (405–560 K), carrier gas humidity (up to 60 vol %), flow velocity (2–6 cm s1), and amount of cesium iodide sorbed on the felt (0.7–65.0 mg). It was found that the oxidative hydrolysis under certain conditions yields CsI CsOH and CsI3 aerosols. Although under the experimental conditions studied the sorption of radionuclides on the metal felt is nearly quantitative (97–99%), escape of the radionuclides from the column upon prolonged passing of steam-gas mixtures cannot be excluded.Translated from Radiokhimiya, Vol. 46, No. 5, 2004, pp. 449–453.Original Russian Text Copyright © 2004 by Kulyukhin, Mikheev, Kamenskaya, Rumer, Konovalova, Novichenko.  相似文献   

13.
The thermal diffusivity of simulated fuels with dissolved fission products was measured by using the laser-flash method in the temperature range from room temperature to 1,473 K. Three kinds of simulated fuels with an equivalent burn-up of 3, 6, and 12 at% were used in the measurement. The thermal diffusivity and the thermal conductivity of the simulated fuels with the dissolved fission products decreased, as the temperature and the equivalent burn-up increased. The thermal conductivities of simulated fuels with equivalent burn-ups of 3, 6, and 12 at% were lower than that of UO2 by 84.70, 67.17, and 44.97% at 300 K and 99.17, 89.88, and 80.56% of UO2 at 1,473 K, respectively. The difference in the thermal conductivity between the simulated fuel and UO2 was large at room temperature, and it decreased as the temperature increased. The thermal resistivity of the simulated fuels increased linearly with temperature up to 1,473 K.  相似文献   

14.
The density of a UO2–ZrO2 melt (atomic ratio U/Zr = 1.528) is experimentally measured by a pycnometric method in the temperature range of 2973–3373 K. The found temperature dependence of density has the form (T) = (7.0 ± 0.01) – (4.5 ± 0.4) × 10–4 × (T – 2973 K), g/cm3. The temperature dependence measured enables one to calculate the values of the density of UO2–ZrO2 melts depending on the temperature and composition for any atomic ratio U/Zr.  相似文献   

15.
The surface tension of liquid Ti90Al6V4 was measured. The samples have been processed containerlessly by electromagnetic levitation, which allows the handling of highly reactive materials and measurements in the undercooled temperature region. The use of digital image processing allows the identification of oscillation modes and calculation of the surface tension from the l = 2 and m = 0, m = 2 oscillation modes. A linear least squares fit to the data showed the following temperature dependence: = 1.389 ± 0.09 – 9.017 × 10–4 ± 5.64 × 10–5(T – 1660°C) [Nm–1]  相似文献   

16.
New values of densities and surface tensions of liquid aluminum obtained in the range 1600 to 2360 K by contactless techniques in neutral gases are reported. Conditions for oxygen-free aluminum are fulfilled which allow determination of the surface tension of aluminum. Extrapolation to the melting point, T m = 933 K, confirms the value of (T = 933K) = 1.05 N m–1.  相似文献   

17.
Differential thermal analysis and x-ray diffraction data indicate that the ZnO B2O3-CuO B2O3 join of the ternary system CuO-B2O3-ZnO is pseudobinary, with eutectic phase relations and a liquid-liquid miscibility gap in the composition range 25–35 mol % CuO.Translated from Neorganicheskie Materialy, Vol. 41, No. 3, 2005, pp. 339–340.Original Russian Text Copyright © 2005 by Kasumova, Bananyarly.This revised version was published online in April 2005 with a corrected cover date.This revised version was published online in April 2005 with a corrected cover date.  相似文献   

18.
The enthalpy and specific heat of a Be2C-Graphite-UC2 composite nuclear fuel material have been measured over the temperature range 298–1980 K using both differential scanning calorimetry and liquid argon vaporization calorimetry. The fuel material measured was developed at Sandia National Laboratories for use in pulsed test reactors. The material is a hot-pressed composite consisting of 40 vol% Be2C, 49.5 vol% graphite, 3.5 vol% UC2, and 7.0 vol% void. The specific heat was measured with the differential scanning calorimeter over the temperature range 298–950 K, while the enthalpy was measured over the range 1185–1980 K with the liquid argon vaporization calorimeter. The normal spectral emittance at a wavelength of 6.5×10–5 cm was also measured over the experimental temperature range. The combined experimental enthalpy data were fit using a spline routine and differentiated to give the specific heat. Comparison of the measured specific heat of the composite to the specific heat calculated by summing the contributions of the individual components indicates that the specific heat of the Be2C component differs significantly from literature values and is approximately 0.56 cal · g–1 · K –1 (2.3×103J · kg–1 · K –1) for temperatures above 1000 K.  相似文献   

19.
A critical analysis of the data for the specific heat from 15 K to 300 K of two samples of 241Am and one sample of 243Am leads to a set of best values. From these the thermodynamic functions are calculated, giving (C p)298 = 25.5 ± 1 J mol–1 K–1 and S 298 = 55.4 ± 2 J mol–1 K–1. The derived Debye characteristic temperature D is estimated as 120 ± 20 K and the electronic specific heat coefficient as 1 ± 1 mJ mol–1 K–2.  相似文献   

20.
Superconducting transition-edge sensors have been used extensively in cryogenic particle detectors, either as thermometers for microcalorimetry or as bolometers for the detection of the prompt phonons resulting from a particle decay in a single crystal absorber. Bolometer action depends upon the energy coupling of the prompt phonons to the bolometer electrons. A study has been made of the electron-phonon coupling for a series of Au-Ti bolometers on a Si substrate and of the use of these bolometers for prompt phonon detection below 1 K. The electron-phonon coupling was found to be proportional to the normalized resistance (R/R n) of the bolometer; R is the bolometer resistance and R n is the normal resistance. When extrapolated to R/R n = 1, this coupling was consistent with /VT 3 = 3 × 109 Wm–3K–4 where is the thermal conductance from the bolometer electrons to the Si phonons and V and T are the volume and transition temperature of the bolometer. The response of the bolometers to heat pulses generated by a thin film heater on the opposite face of a Si single crystal were similar to that generally seen above 1 K, apart from a delay time constant that varied from 0.5 to 1.3 µs as the transition temperature decreased from 600 to 200 mK. This delay time constant is attributed to the thermal equilibrium time of normal regions of the bolometer.  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