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 375 毫秒
1.
A mass spectrometric Knudsen effusion method has been used to study the vaporization behaviors of three sodium borosilicate glasses with the compositions of 1Na2O-1B2O3-3SiO2(Glass-1), 1Na2O-1.5B2O3-3SiO2(Glass-2) and 1.5Na2O-1B2O3-3SiO2(Glass-3) in the temperature range 915–1172 K. Vapor species of NaBO2(g) and Na2(BO2)2(g) have been identified over Glass-1 and Glass-2, in which the mole ratio of Na2O to B2O3 is equal to and less than unity, respectively. Over Glass-3, in which the ratio is larger than unity, vapor species of Na(g) has been observed in addition to NaBO2(g) and Na2(BO2)2(g), and the total vapor pressure over Glass-3 is much higher as compared with those over Glass-1 and Glass-2. From the measured partial pressures of Na(g), NaBO2(g) and Na2(BO2)2(g), the activity coefficient of NaBO2 in Glass-1, enthalpies of vaporization, enthalpies of formation and dissociation energies for the vapors have been determined.  相似文献   

2.
The vaporization of Li2SiO3(c/1) has been studied by the mass spectrometric Knudsen effusion method. The vaporization process has been found to be incongruent. Partial pressures of Li(g), LiO(g), Li2O(g), SiO(g) and Li2SiO3(g) over Li2SiO3(c/1) have been determined in the temperature range 1166–1762 K. Partial pressures of O2(g) have also been calculated from the reaction Li2SiO3(1) = 2 Li(g) + SiO(g) + O2(g). The enthalpies of formation and the atomization energies for LiO(g) and Li2O(g) have been evaluated from the partial pressures to be ΔHof0(LiO,g) = (84.5 ± 12.8)kJ/mol, ΔHof0(Li2O, g) = (?148.1 ± 15.8)kJ/mol, D0o(LiO) = (321.4 ± 12.8)kJ/mol and D0o(Li2O) = (713.2 ± 15.8)kJ/mol, respectively. The value of D0o(Li2O) is somewhat greater than twice that of D0o(LiO).  相似文献   

3.
The difference in electrochemical corrosion potential of stainless steel exposed to high temperature pure water containing hydrogen peroxide (H2O2) and oxygen (O2)is caused by differences in chemical form of oxide films. In order to identify differences in oxide film structures on stainless steel after exposure to H2O2 and O2 environments, characteristics of the oxide films have been examined by multilateral surface analyses, e.g., X-ray diffraction (XRD), Rutherford back scattering spectroscopy (RBS), secondary ion mass spectroscopy (SIMS) and X-ray photoelectron spectroscopy (XPS). Preliminary characterization results of oxide films confirmed that the oxide film formed under the H2O2 environment consists mainly of hematite (α-Fe2O2), while that under the O2 environment consists of magnetite (Fe3O4). Furthermore oxidation at the very surface of the film is much more enhanced under the H2O2 environment than that under the O2 environment. It was speculated that metal hydroxide plays an important role in oxidation of stainless steel in the presence of H2O2. The difference in electric resistance of oxide film causes the difference in anodic polarization properties. It is recommended that several anodic polarization curves for specimens with differently oxidized films should be prepared to calculate ECP based on the Evans diagram.  相似文献   

4.
A mass spectrometric Knudsen effusion study of the vaporization of LiCrO2 in the temperature range of 1673–1873 K has shown the following: (1) The major vapor species over solid LiCrO2 are Li(g), Cr(g), CrO(g), CrO2(g) and LiCrO2(g). (2) The vaporization process involves a sublimation reaction, LiCrO2(s) = LiCrO2(g), and a dissociation reaction, LiCrO2(s) = 12Cr2O3(s) + Li(g) + 14O2(g). (3) The standard enthalpy of solid LiCrO2 at 298 K is derived to be (?935 ± 21) and (?966 ± 18) kJ/mol from the 2nd and 3rd law treatments, respectively.  相似文献   

5.
The reaction between fission products Mo(VI) and Zr(IV) in hot nitric acid solution during the spent fuel dissolution process leads to the formation of hydrated zirconium molybdate precipitate. This solid can include other metallic cations at the 4th oxidation state (+ IV) inside the crystal such as Pu(IV).

Precipitations in the inactive surrogate system of mixed zirconium-cerium molybdates Mo-Zr(IV)-Ce(IV) were performed all over the cerium molar percentage scale from 0 to 100%, Ce(IV) being used as a surrogate of Pu(IV). The crystal structures were identified from powder X-ray diffraction patterns and complementary investigations by SEM and elemental analyses were carried out. The solid composition pointed out two phases that are ZrMo2O7(OH)2(H2O)2 and Ce3Mo6O24(H2O)4. Each of these compounds was proved being a solid solution within which the ZrIV and CeIV atoms can substitute each other.

Moreover, the influence of the cerium molar fraction on the precipitate solubility was investigated and showed a strong evolution of solubility not only with the nature of the precipitate and the Ce content, but also with the nitric acid concentration.  相似文献   

6.
The [Cu(L)2] · H2O (1), [Co(L)2] · 2H2O (2) and [Zn(L)2] · H2O (3) complexes with leucine (L) as ligand, were synthesized in water solution and analyzed by physical-chemical and spectroscopic means (UV-VIS, FT-IR, ESR). The comparative analysis of the IR spectra for the ligand and the complexes indicate the coordination of the metallic centre to the carboxylic oxygen atom and the nitrogen atom of the amino group due to the shift of the νs(CO) and νs(NH) stretching vibrations. Spectral UV-VIS and ESR data confirmed the covalent metal-ligand bonds, the pseudotetrahedral symmetry around the copper and zinc ions and the octahedral environment for the cobalt ion.  相似文献   

7.
Abstract

The dissolution behavior of Cu metal in hydrogen peroxide (H2O2)-ethylenediaminetetraacetic acid (EDTA) solution was studied through both electrochemical and chemical tests. In the electrochemical experiments, the open circuit potential (OCP) of Cu electrode at low H2O2concentration below 0.5wt% was determined by Cu anodic reaction and H2O2 cathodic reaction. On the other hand, the OCP in higher concentration above 2 w/t% of H2O2 approached to 70 mV vs. SCE, since the redox reaction of H2O2 was so dominant reaction that the Cu anodic reaction was masked. The anodic polarization curve of Cu around the OCP of 70 mV in the range between -200 and 200 mV in the solution without H2O2 showed the maximum Cu dissolution reaction front, while those of both carbon steel and Inconel 600 in the same potential range were situated in the passivation region. The anodic polarization behaviors of Cu in the H2O2 concentrations above 1 wt% were examined by a potentiostatic method along with a weight loss measurement. The anodic Cu dissolution current estimated from the weight loss measurement in the solution containing more than 1 wt% of H2O2 was higher than that predicted by the polarization curve of Cu in the H2O2-free solution. The dissolution reaction rate of Cu in the H2O2 solution linearly increased with H2O2 concentration. This result indicated that the dissolution reaction rate of Cu in the presence of H2O2 was enhanced by a reaction between Cu and H2O2, resulting in the shift of the polarization curves depending on H2O2 concentration. The experimental correlation equation based on the above results described well the overall dissolution reaction rate of Cu in H2O2 concentration ranges below 1 wt% and above 2 wt% in the EDTA solution.  相似文献   

8.
The thermodynamic stability of rubidium thorate, Rb2ThO3(s), was determined from vaporization studies using the Knudsen effusion forward collection technique. Rb2ThO3(s) vaporized incongruently and predominantly as Rb2ThO3(s)=ThO2(s) + 2Rb(g) + 1/2 O2(g). The equilibrium constant K=pRb2·pO21/2 was evaluated from the measurement of the effusive flux due to Rb vapor species under the oxygen potential governed by the stoichiometric loss of the chemical component Rb2O from the thorate phase. The Gibbs energy of formation of Rb2ThO3 derived from the measurement and other auxiliary data could be given by the equation, ΔfG°(Rb2ThO3,s)=−1794.7+0.42T ± .  相似文献   

9.
Effective atomic numbers for (PbO and Na2B4O710H2O) and (UO2(NO3)2, and Na2B4O710H2O) mixtures against changing contents of PbO, Na2B4O710H2O, and UO2(NO3)2 were measured in the X-ray energy range from 25.0 to 58.0 keV. The gamma rays emitted by a 241Am annular source have been sent on the absorbers which emits their characteristic X-rays to be used in transmission arrangement. The X-rays were counted by a Si(Li) detector with a resolution of 146 eV at 5.90 keV. The changing compositions of the compounds were assigned to be 0, 0.167, 0.333, 0.500, 0.666, 0.833 and total masses of the mixtures were adjusted to be identical. Also, the total effective atomic numbers of each mixture were estimated by using the mixture rule. The measured values were compared with estimated values for the mixtures.  相似文献   

10.
Photoluminescence (PL) spectrum, in conjunction with X-ray photoelectron spectroscopy (XPS) is used to evaluate the surface damage of GaN layer by highly-charged Xeq+ (18 ? q ? 30), Arq+ (6 ? q ? 16) and Pbq+ (q = 25,35) ions. The intensity of PL emission of GaN layer, including near band-edge peak and yellow luminescence, decreases with increasing fluence and charge state of the incident ions. Finally the PL emission is completely quenched after irradiation to high fluences at high charge state. A new peak at 450 nm appeared in PL spectra of the specimens irradiated with Xe18+, Ar6+ and Ar11+, indicating that radioactive recombination within donor-acceptor pairs (DAPs) during irradiation. After irradiation, XPS spectra show N deficient or Ga rich on GaN surface and XPS spectra of Ga3d core levels indicate spectral peak evidently shifts from a Ga-N to Ga-Ga and Ga-O bond. The relative content of Ga-N bond decreases and the content of Ga-Ga bond increases with the increase of ion fluence and ion charge state. The binding energy of Ga3d5/2 electron corresponding to Ga-Ga bond of the irradiated GaN film is found to be smaller than that of metallic Gallium (Ga0), which can be attributed to irradiation damage.  相似文献   

11.
(Y, La)3(Fe, Ga)5O12 epitaxial garnet films on (111) Gd3Ga5O12 substrates irradiated with 238U ions of 1.4 MeV/u specific energy in the dose range 1010 cm?2 to 3 × 1011 cm?2 were measured by means of Rutherford backscattering and double-crystal X-ray diffraction before and after thermal annealing in oxygen. The nuclear track diameter of 10 nm confining a cylindrical volume of highly disordered material caused by each ion impact has been deduced from the comparison of the backscattering spectra of the irradiated and unirradiated film areas. The fraction of randomly backscattered ions due to the irradiation-induced damage as well as the lattice expansion perpendicular to the crystal surface caused by irradiation-induced lateral compressive stress are proportional to the ion dose. After thermal annealing the comparison of the almost identical backscattering yield of the irradiated areas and the unirradiated film regions demonstrates a nearly perfect recrystallization of the damaged track volumes.  相似文献   

12.
A Focused Ion Beam (FIB) has been used to implant micrometer-sized areas of polycrystalline anatase TiO2 thin films with Ga+ ions using fluencies from 1015 to 1017 ions/cm2. The evolution of the surface morphology was studied by scanning electron microscopy (SEM) and atomic force microscopy (AFM). In addition, the chemical modifications of the surface were followed by X-ray photoelectron spectroscopy (XPS). The implanted areas show a noticeable change in surface morphology as compared to the as-deposited surface. The surface loses its grainy morphology to gradually become a smooth surface with a RMS roughness of less than 1 nm for the highest ion fluence used. The surface recession or depth of the irradiated area increases with ion fluence, but the rate with which the depth increases changes at around 5 × 1016 ions/cm2. Comparison with implantation of a pre-irradiated surface indicates that the initial surface morphology may have a large effect on the surface recession rate. Detailed analysis of the XPS spectra shows that the oxidation state of Ti and O apparently does not change, whereas the implanted gallium exists in an oxidation state related to Ga2O3.  相似文献   

13.
The vapor pressures over UO2.000 and (U1?yNby)O2+x (y = 0.01, 0.05, x = 0.000–0.022) were measured by the mass-spectrometric method in the temperature range 2025–2343 K. The main gas species over UO2.000 were observed to be UO3(g) and UO2(g) and those over (U1?yNby)O2+x were NbO2(g), NbO(g), UO3(g) and UO2(g). The partial vapor pressures of almost all gas species over (U1?yNby)O2+x increased with increasing O/M (M = U + Nb) ratio. With increasing Nb content in (U1?yNby)O2.000, the partial vapor pressures of UO2(g) and UO3(g) decreased and those of NbO(g) and NbO2(g) increased. The congruently vaporizing composition in the (U1?yNby)O2+x phase was estimated to be (U0.985±0.005Nb0.015±0.005)O2.000 from the compositional dependence of the total vapor pressures. The partial molar enthalpy and entropy of oxygen of (U1?yNby)O2+x calculated from the partial pressures of gaseous species NbO2(g) and NbO(g) were in fairly good agreement with those previously obtained by the present authors with a thermobalance.  相似文献   

14.
The depositionof Ni and Co ions on a heated surface of simulating fuel rods has been studied in water at atmospheric and 70 atm pressures during nucleate boiling.

The effects of various factors, including heat flux of the heated surface, concentration of coexisting iron oxide (αFe2O3 and concentration of Ni and Co ions, on their deposition rate have been investigated. The model for iron oxide deposition which is based on microlayer evaporation and drying out phenomena in the nucleate boiling bubble was shown to be applicable to the deposition of Ni and Co ions. That is, dW/dt=K.Q.C/L, where dW/dt is the deposition rate, K the deposition rate coefficient, Q the heat flux, C the ion concentration, and L the latent heat of vaporization. The K value of Ni ion is about 0.1 and independent of iron oxide concentration. On the hand, the K value of Co ion increases with iron oxide concentration and seems to approach that of iron oxide concentration and seems to approach that of iron oxide (0.3). The Co ion deposited with iron oxide forms Co ferrite. Solubility of Co ferrite is small compared with that of Co deposits without iron oxide (CoO or Co(OH)2). The increase in the K value of Co ion with iron oxide concentration is attributed to the change in chemical form of Co deposits into more stable species not favoring Co release.  相似文献   

15.
The results of present paper have shown that sputtering of yttrium iron garnet (Y3Fe5O12) under swift heavy ions in the electronic energy loss regime is non-stoichiometric. Here we are presenting additional experimental results for gadolinium gallium garnet (Gd3Ga5O12) as target. The irradiations were performed with different ions (50Cr (589 MeV), 86Kr (195 MeV) and 181Ta (400 MeV)) impinging perpendicularly to the surface. As earlier, the sputtering yield was determined by collecting the emitted gadolinium and gallium atoms on a thin aluminium foil, placed upstream above the target and analyzing the Al catcher by Rutherford backscattering. Also for Gd3Ga5O12, the emission of Gd and Ga is non-stoichiometric. Sputtering appears above a critical electronic stopping power of Sth = 11.6 ± 1.5 keV/nm, which is larger than the threshold for track formation, in agreement with other amorphisable materials. In addition, the angular distribution of the sputtered species was measured for Y3Fe5O12 and Gd3Ga5O12 using 200 MeV Au ions impinging the surface at 20° relatively to the surface. For the two garnets the ratio of Y/Fe (and Gd/Ga) varies with the angle of emitted species and the stoichiometry seems to be preserved only for an emission perpendicular to the surface.  相似文献   

16.
The precipitations of thorium and uranium(VI) sulfito complex ions with hexammine cobalt(III) chloride as the precipitant have been studied.

The orange-colored uranium(VI) precipitate obtained is [Co(NH3)6]4[UO2(SO3)3]322H2O, which is in the form of square bipyramid, about 4 μm across in a cubic symmetry of the diamond type with a=10.40Å It decomposes to an oxide mixture of Co3O4 and U3O8 above 850°C in the air through a sulfate mixture of CoSo4 and UO2SO4.

Composition of the thorium precipitate varies with the precipitation conditions. Therefore, it is considered that the thorium precipitate contains thorium hydroxide and basic thorium sulfite.  相似文献   

17.
18.
An analysis is made of the chemical heats liberated from a palladium-deuterium electrochemical cell operating inside a calorimeter. It is important, in such an analysis, to carefully identify the chemical and electrochemical sources of heat, before any “excess heats” can be ascribed to non-chemical reactions.(1) The calorimeter measures the enthalpy (ΔH r ) of the reaction; whereas, the electrochemical voltage of the cell reflects the free energy (ΔG r ) of the reaction within the Pd-D electrolysis cell. The heat energy from the calorimeter cell therefore doesnot equal the electrical energy supplied to the cell, as might initially be expected. The magnitudes of the differing calorimetric and electrochemical energies were found to be related through the “thermoneutral potential” (ξH) of the electrochemical reaction. The chemical heat theoretically expected from the calorimeter is given by (1) I(ξLH), the cell current (I), multiplied by the difference between the operating cell voltage (ξL) and the thermoneutral potential (ξH), rather than (2) IξL, the electrical input power. This was verified empirically using a freon vaporization calorimeter, which operates on the principle of accurate measurement of the vaporization rate of liquid freon which completely surrounds the electrochemical cell. The calorimetrically-measured heats observed from a Pt/D2O, 0.1M LiOD/Pd electrochemical cell were within 2% of the thermoneutral potential predicted value, I(ξLH); but were found to be 15–30% less than the electrical work supplied to the cell, IξL. Measurements of D2O consumed by the cell reactions also verified that essentially no significant recombination of D2 and O2 gases occurred within the cell. No “excess heats” were observed from this Pd cell during the 36 days of its electrolytic operation. Likewise, no increase in the neutron flux around the cell was found, using three3He radiation detectors.  相似文献   

19.
20.
To predict the fundamental phase relationships in the solidified core melt of the Fukushima Daiichi Nuclear Power Plant, solidified melt samples of the various core materials [B4C, stainless steel, Zr, ZrO2, (U,Zr)O2] were prepared by arc melting. Phases and compositions in the samples were determined by means of X-ray diffraction, microscopy, and elemental analysis. With various compositions, the only oxide phase formed was (U,Zr)O2. After annealing, the stable metallic phases were an Fe-Cr-Ni alloy and an Fe2Zr-type (Fe,Cr,Ni)2(Zr,U) intermetallic compound. The borides, ZrB2 and Fe2B-type (Fe,Cr,Ni)2B, were solidified in the metallic part. Annealing at 1773 K under an oxidizing atmosphere (Ar-0.1%O2) resulted in the oxidation of U and Zr in the alloy and in ZrB2, and consequently the (Fe,Cr,Ni)2B and Fe-Cr-Ni alloy became dominant in the metallic part. The experimental phase relationships in the metallic part agreed reasonably with the thermodynamic evaluation of equilibrium phases in a simplified B4C–Fe–Zr system. The metallic Zr content in the melt was found to be a key factor in determining the phase relationships. As a basic mechanical property, the microhardness of each phase was measured. The borides, especially ZrB2, showed notably higher hardness than any other oxide or metallic phases.  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