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 359 毫秒
1.
The vapor pressures of CdI2 and Cs2CdI4 were measured below and above their melting points, employing the transpiration technique. The standard Gibbs energy of formation ΔfG° of Cs2CdI4, derived from the partial pressure of CdI2 in the vapor phase above and below the melting point of the compound could be represented by the equations ΔfG°Cs2CdI4 (±6.7) kJ mol−1=−1026.9+0.270 T (643 K≤T≤693 K) and ΔfG°{Cs2CdI4} (±6.6) kJ mol−1=−1001.8+0.233 T (713 K≤T≤749 K) respectively. The enthalpy of fusion of the title compound derived from these equations was found to be 25.1±10.0 kJ mol−1 compared to 36.7 kJ mol−1 reported in the literature from differential scanning calorimetry (DSC). The standard enthalpy of formation ΔfH°298.15 for Cs2CdI4 evaluated from these measurements was found to be −918.0±11.7 kJ mol−1, in good agreement with the values −920.3±1.4 and −917.7±1.5 kJ mol−1 reported in the literature from two independent calorimetric studies.  相似文献   

2.
The enthalpy of γ-LiAlO2 was measured between 403 and 1673 K by isothermal drop calorimetry. The smoothed enthalpy curve between 298 and 1700 K results in H0(T) − H0(298 K)=−37 396 + 93.143 · T + 0.00557 · T2 + 2 725 221 · T−1 J/mol. The standard deviation is 2.2%. The heat capacity was derived by differentiation of the enthalpy curve. The value extrapolated to 298 K is Cp,298=(65.8 ± 2.0) J/K mol.  相似文献   

3.
The thermal conductivity, λ of a saturated vapor over UO1.96 is calculated in the temperature range 3000–6000 K. The calculation shows that the contribution to λ from the transport of reaction enthalpy dominates all other contributions. All possible reactions of the gaseous species UO3, UO2, UO, U, O, and O2 are included in the calculation. We fit the total thermal conductivity to the empirical equation λ = exp(a+ b/T+cT+dT2 + eT3), with λ in cal/(cm s K), T in kelvins, a = 268.90, B = − 3.1919 × 105, C = −8.9673 × 10−2, d = 1.2861 × 10−5, and E = −6.7917 × 10−10.  相似文献   

4.
The vaporization of Li2TiO3(s) has been investigated by the mass spectrometric Knudsen effusion method. Partial pressures of Li(g), LiO(g), Li2O(g), Li3O(g) and O2(g) over Li2TiO3(s) have been obtained in the temperature range 1180–1628 K. When the vaporization of Li2TiO3(s) proceeds, the content of Li2O in the Li2TiO3(s) sample decreases. The phase of the sample is a disordered Li2TiO3 solid solution above 1486 K. The enthalpies of formation and the atomization energies for LiO(g) and Li3O(g) have been evaluated from the partial pressures to be ΔHof0(LiO, g) = 65.4 ± 17.4 kJ/mol, ΔHof0(Li3O, g) = − 207.5 ± 56.6 kJ/mol, Do0(LiO) = 340.5 ± 17.4 kJ/mol and Do0(Li3O) = 931.6 ± 56.6 kJ/mol, respectively.  相似文献   

5.
Brannerite, ideally MTi2O6, (M=actinides, lanthanides and Ca) occurs in titanate-based ceramics proposed for the immobilization of plutonium. Standard enthalpies of formation, ΔH0f at 298 K, for three brannerite compositions (kJ/mol): CeTi2O6 (−2948.8 ± 4.3), U0.97Ti2.03O6 (−2977.9 ± 3.5) and ThTi2O6 (−3096.5 ± 4.3) were determined by high temperature oxide melt drop solution calorimetry at 975 K using 3Na2O · 4MoO3 solvent. The enthalpies of formation were also calculated from an oxide phase assemblage (ΔH0f-ox at 298 K): MO2 + 2TiO2=MTi2O6. Only UTi2O6 is energetically stable with respect to an oxide assemblage: U0.97Ti2.03O6H0f-ox=−7.7±2.8 kJ/mol). Both CeTi2O6 and ThTi2O6 are higher in enthalpy with respect to their oxide assemblages with (ΔH0f-ox=+29.4±3.6 kJ/mol) and (ΔH0f-ox=+19.4±1.6 kJ/mol) respectively. Thus, Ce- and Th-brannerite are entropy stabilized and are thermodynamically stable only at high temperature.  相似文献   

6.
The sodium potential in the test electrode (a) Pt,O2,Na2ZrO3,ZrO2 was measured by using the emf technique employing Na-β-alumina as the solid electrolyte in conjunction with (b) Pt,O2,Al2O3,NaAl11O17, (c) Pt,O2,Na2MoO4,Na2Mo2O7 and (d) Pt,Na2CO3,CO2,O2 as the reference electrodes over the ranges 880–1045, 700–800 and 850–940 K, respectively. The emf results between electrodes (b) and (c) were utilized for internal consistency checks. From the results on cells formed between (a) and (b) and those on (a) and (c), the standard Gibbs energy of formation, ΔfGo (kJ/mol) of Na2ZrO3 was determined to be −1699.4+0.3652T (K) valid over the temperature range 700–1045 K. The break in the emf data at 1045 K was corroborated by independent TG/DTA measurements carried out on Na2ZrO3 which exhibited an endotherm at 1055 K indicative of a phase transition in Na2ZrO3.  相似文献   

7.
Energy and angular distributions of Cr+ sputtered from stainless steel by 1.6 × 10−15 J (10 keV) H+3 are reported as a function of angle of incidence. For more normal incidence, the peak in the energy distribution occurs in the vicinity of 3.2 × 10−19 J (2 eV), the average energy is approximately 1.12 × 10−18 J (7 eV), and the angular distribution is close to cosine. Toward glancing incidence, the peak energy increases to ˜6.4 × 10−19 J (4 eV), the average energy increases to ˜1.28 × 10−18 J (8.0 eV), and the angular distribution shows a distinct maximum in the forward direction. These results are discussed in terms of the increasing role of surface recoils in the sputtering mechanism at glancing incidence.  相似文献   

8.
The EMF of the following galvanic cells,
(render)
Kanthal,Re,Pb,PbOCSZO2 (1 atm.),Pt
(render)
Kanthal,Re,Pb,PbOCSZO2(1 atm.),RuO2,Pt
were measured as a function of temperature. With O2 (1 atm.), RuO2 as the reference electrode, measurements were possible at low temperatures close to the melting point of Pb. Standard Gibbs energy of formation, ΔfG0mβ-PbO was calculated from the emf measurements made over a wide range of temperature (612–1111 K) and is given by the expression: ΔfG0mβ-PbO±0.10 kJ=−218.98+0.09963T. A third law treatment of the data yielded a value of −218.08 ± 0.07 kJ mol−1 for the enthalpy of formation of PbO(s) at 298.15 K, ΔfH0mβ-PbO which is in excellent agreement with second law estimate of −218.07 ± 0.07 kJ mol−1.  相似文献   

9.
A knowledge of the threshold oxygen level in liquid sodium necessary for the formation of NaCrO2 in sodium-steel systems is useful in the operation of fast breeder reactors. There is considerable discrepancy in the data reported in the literature. In order to resolve this, the problem was approached from two sides. Direct measurement of oxygen potential in the Na(l)-Cr(s)-NaCrO2(s) phase field using the galvanic cell In, In2O3/YDT/Na, Cr, NaCrO2 yielded: o2 = −800847 + 147.85 T J/mol O2 (657–825 K). Knudsen cell-mass spectrometric measurements were carried out in the phase field NaCrO2(s)-Cr2O3(s)-Cr(s) to obtain the Gibbs energy of formation of NaCrO2 as: ΔGof,T(NaCrO2) = −870773 + 193.171 T J/mol (825–1025 K). The threshold oxygen levels deduced from Gof,T (NaCrO2) data were an order of magnitude lower than the directly measured values. The difference between the two sets of data as well as differing experimental observations from operating liquid sodium systems are explained on the basis of the influence of dissolved carbon.  相似文献   

10.
The diffusion behavior of tritium in UO2 was studied. Two methods were adopted for the introduction of tntium into UO2: one via ternary fission of 235U and the other via thermal doping. In the former, the diffusion constants decreased with increase in sample weight. The diffusion constants obtained from the pellet with the same specification (9 mm in diameter, 5 mm high) were Dbulk = 3.03 × 10−3(+0.369−0.003) exp[−163±43(kJ/mol)/RT](cm2/s) for fission-created tritium and Dbulk = 0.15(+ 0.94−0.13) exp[−76±13 (kJ/mol)/RT](cm2/s) for thermally-doped tritium. The difference of the diffusion constants between two systems was discussed in terms of the effects associated with the recoil processes of energetic tritium.  相似文献   

11.
Isothermal sections of the phase diagram of the system MgO–Al2O3–PuO2 at various temperatures were calculated using sublattice models. The results show that below 2133 K no liquid occurs in the system. Above 2133 K liquid starts to form at the Al2O3–PuO2 side. The phase diagram of the pseudo-binary system PuO2–MgAl2O4 was also obtained from an isopleth Tx calculation.  相似文献   

12.
Creep experiments were conducted on nearly stoichiometric UO2 helical springs from 1000 to 1600°C and 2.1 to 80 MPa. Entirely transient behaviour was measured in all experiments with the plastic strain,ε = (Aσ/d1.5) exp(−Q/RT)tm, where A is a constant that depends on purity, d is the grain size, σ is the applied stress, Q is the apparent activation energy, t is the time, m is a constant, and the other terms have their usual meaning. At T > 1200°C, Q 100 kcal mol−1, but at T < 1200°C, Q increased dramatically and became strain dependent. The value of m for most experiments was 0.8, but at σ > 48 MPa, m decreased, and for d < 10 μm, it increased. Amorphous or glassy grain boundary phases were observed by transmission electron microscopy in all specimens: specimens containing the largest concentrations of Fe and Si sometimes had anomalously high creep rates. The phases existed as discontinuous, lenticular bodies on grain faces and a continuous network along triple grain junctions. Some instances of precipitation of UO2 from the phase were observed. At T > 1200°C, glassy phases may accelerate Coble creep by providing short circuit diffusion paths along the grain boundaries or may accelerate superplastic deformation by diffusion along the continuous glassy phase triple line junctions. At low temperatures the glassy phase appears to control grain-boundary sliding.  相似文献   

13.
The Vickers micro-hardness (HV) was measured by an indentation technique of simulated ZrO2-based Inert Matrix Fuel (IMF) material with a composition of Er0.07Y0.10Ce0.15Zr0.68O1.915 in two different densities on sintered specimens and specimens thermally shocked with the quenching temperature differences (ΔTs) between 473 and 1673 K and compared with those of simulated MOX, namely, U0.92Ce0.08O2. The HV values obtained for two IMF materials were found higher, ranging from 6.37 GPa to about 7.84 GPa, depending on ΔT and the sintered density, than those obtained for the simulated MOX which are quasi-constant in the same range of ΔT with a mean value of 6.37 GPa. The fracture toughness (KIC) was calculated from the measured HV and the crack length, and it was found to exhibit a slight increase with increasing ΔT, ranging between 1.4 and 2.0 MPa m1/2, while that of simulated MOX specimen is ranging between 0.8 and 1.1 MPa m1/2. The thermally shocked specimens were observed with an optical microscope and analyzed in terms of microstructural changes and cracking patterns.  相似文献   

14.
Polycrystalline pellets of the rare-earth sesquioxide Dy2O3 with cubic C-type rare-earth structure were irradiated with 300 keV Kr2+ ions at fluences up to 5 × 1020 Kr/m2 at cryogenic temperature. Irradiation-induced microstructural evolution is characterized using grazing incidence X-ray diffraction (GIXRD) and transmission electron microscopy (TEM). In previous work, we found a phase transformation from a cubic, C-type to a monoclinic, B-type (C2/m) rare-earth structure in Dy2O3 during Kr2+ ion irradiation at a fluence of less than 1 × 1020 Kr/m2. In this study, we find that the crystal structure of the top and middle regions of the implanted layer transform to a hexagonal, H-type (P63/mmc) rare-earth structure when the irradiation fluence is increased to 5 × 1020 Kr/m2; the bottom of the implanted layer, on the other hand, remains in a monoclinic phase. The irradiation dose dependence of the C-to-B-to-H phase transformation observed in Dy2O3 appears to be closely related to the temperature and pressure dependence of the phases observed in the phase diagram. These transformations are also accompanied by a decrease in molecular volume (or density increase) of approximately 9% and 8%, respectively, which is an unusual radiation damage behavior.  相似文献   

15.
Previous work on diffusion in inert-gas bombarded Al2O3 has revealed the presence of four diffusion processes, of which two take place well below the temperatures for self-diffussion, one agrees with self-diffusion, and one occurs at temperatures well above those for self-diffusion. The present work serves to explore in greater detail the two low-temperature processes. It is shown that the first, which is found in -Al2O3, Al(OH)3, and γ-Al2O3beginning at about 100° C, is consistent with a range of ΔH's of 28 to 50 kcal/mole. The mechanism of the process is hinted at by the fact that it overlaps the temperatures both for Al(OH)3decomposition and for point-defect motion in -Al2O3; the correlation with point defects is believed, however, to be the more significant. The second process, which is found only in -Al2O3 beginning at 500–650° C, implies an essentially single ΔH lying between 69 and 79 kcal/mole. It was suggested previously by Matzke and Whitton on the basis of electron diffraction that the process could be attributed to the amorphous-crystalline transition of -Al2O3. Further aspects of low-temperature diffusion in Al2O3 were revealed by comparing autoradiographs of specimens of -A2O3which were bombarded to various doses and then either heated to 850° C or immersed in unheated 12N NaOH. Thus regions exposed to a high dose and which would be expected to be amorphous, had an increased sticking factor, a greater tendency to lose gas during heating, and an enhanced chemical reactivity.  相似文献   

16.
Interaction processes resulting from the transit of incident 2–30 keV H+, H+2 and H+3 through 1.2 to 2.5 μg cm−2 carbon foils are investigated by examining the charge state and angular scatter distributions of atomic and molecular species that exit the foils. A comparison of the scatter distributions of exiting H+2 and H0 from incident H+2 and H+3 show that the atomic components of transmitted molecules scatter independently from foil atoms. For a given foil thickness, the measured fractions of H+2 from incident H+2 and H+3 are inversely proportional to the square of the angular scatter half-width.  相似文献   

17.
The deposition of high-quality high-Tc superconducting films on silicon wafers for future hybrid electronic devices is strongly hampered by the interdiffusion between films and substrate. This effect degrades the superconducting properties seriously and is a strong function of temperature. Since high processing temperatures are inevitable for good films, suitable buffer layers are needed to reduce the interdiffusion. We have investigated the combinations ZrO2/Si(100), BaF2/Si(100), and noble-metal/TiN/Si(100) at temperatures up to 780°C in oxidizing ambient. YBa2Cu3O7−x films have been deposited onto the buffer layers by laser ablation. Thereafter the interfaces have been analyzed by Rutherford backscattering. So far only ZrO2 has demonstrated sufficient stability to serve as a buffer layer for the laser-ablated YBa2Cu3O7−x films. All other combinations suffer from interdiffusion or oxidation.  相似文献   

18.
19.
We summarize the diametral creep results obtained in the MR reactor of the Kurchatov Institute of Atomic Energy on zirconium-2.5 wt% niobium pressure tubes of the type used in RBMK-1000 power reactors. The experiments that lasted up to 30 000 h cover a temperature range of 270 to 350°C, neutron fluxes between 0.6 and 4.0 ×1013 n/cm2 · s (E > 1 MeV) and stresses of up to 16 kgf/mm2. Diametral strains of up to 4.8% have been measured. In-reactor creep results have been analyzed in terms of thermal and irradiation creep components assuming them to be additive. The thermal creep rate is given by a relationship of the type εth = A1 exp [(A2 + A t) T] and the irradiation component by εrad = Atø(TA5), where T = temperature, σt = hoop stress, ø = neutron flux and a1 to A5 are constants. Irradiation growth experiments carried out at 280° C on specimens machined from pressure tubes showed a non-linear dependence of growth strain on neutron fluence up to neutron fluences of 5 × 1020 n/cm2. The significance of these results to the elongation of RBMK reactor pressure tubes is discussed.  相似文献   

20.
Kinetics of the carbothermic synthesis of UN from UO2 in an NH3 stream and a mixed 75% H2 + 25% N2 stream were studied in the temperature range of 1400–1600°C by X-ray analysis and weight change measurement of the sample. The weight change was divided into two parts; i.e. weight loss due to carbothermic reduction of UO2 and weight loss due to removal of carbon by hydrogen. The former followed the first-order rate equation −1n(1 − 0) = k0t, and the latter the rate equation of phase boundary reaction 1 − (1 − c)1/3 = kct. The apparent activation energy of the former was in the range of 320–380 kJ/mol. The value of the latter in an NH3 stream was 175–185 kJ/mol, which was smaller than that in a mixed 75% H2 + 25% N2stream (285 kJ/mol). In this method, the rate of the removal of carbon by hydrogen determines that of the formation of high purity UN.  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