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.
The enthalpy stability of the LaCl 4 ? and LuCl 4 ? ions is assessed using high-temperature mass spectrometry. The enthalpy of Cl? detachment is determined to be ΔrH0(298.15 K) = 332 ± 10 kJ/mol for LaCl 4 ? and 359 ± 10 kJ/mol for LuCl 4 ? .  相似文献   

2.
The absorption spectra of the NpO 2 + (5f 2) ion were examined in the region of the 3H 53 H 4 magnetic dipole transition (1530–1760 nm) for series of melts with the UO 2 2+ concentration varied in the opposite directions: (1) NaCl-2CsCl eutectic melt with growing additions of the Cs2UO2Cl4 complex salt and (2) Cs2UO2Cl4 melt with growing additions of the NaCl-2CsCl mixture. Measurements of the integrated intensities of the bands belonging to the NpO 2 + ·UO 2 2+ complex and unbound NpO 2 + throughout the UO 2 2+ concentration range examined (up to 4.4 M in neat Cs2UO2Cl4 melt) and processing of the data obtained in terms of the mass action law showed that the formation-decomposition reaction of the cation-cation complex can be described adequately only using the equation of reaction in the form NpO2Cl 4 3? + UO2Cl 4 2? ? {Cl4ONpO?UO2Cl3}4? = Cl? (with the equilibrium constant of 1.3±0.1). Thus, the formation of the cationcation complex should be treated as replacement of chloride ion in the equatorial plane of uranyl(VI) by neptunyl(V), rather than as simple addition of UO 2 2+ to NpO 2 + . The reverse reaction, decomposition of the cation-cation complex, consists essentially in replacement of neptunyl(V) by chloride ion.  相似文献   

3.
The mass spectrum of the products of arc discharge in helium between graphite electrodes has been studied for various values of the gas flow rate. As the gas flow rate increases, the intensity of C60±, C70±, C84± and C90± fullerene peaks increases and that of the C2 and C3+ cluster radicals decreases, but the total decay in radicals amounts to only 21% of the total growth of fullerenes. From this it follows that a contribution to the formation of fullerenes from the neutral clusters (which are taken into account for the first time) significantly exceeds the contribution due to small radical species.  相似文献   

4.
Results of experiments on determination of the energy of 213Bi α-particles for the existing lines with the intensities of 1.94 and 0.15% per decay are reported. The results were obtained by semiconductor α-ray spectrometry by comparison with the α-particle energies for the major lines of 221Fr and 213Po, known with a lower uncertainty. The experiments were designed so as to minimize and, at the same time, take into account factors affecting the deviations of the peak maxima from the true values. Special emphasis is made on the decay of recoil nuclei on the detector surface. To eliminate this effect, samples containing highly nonequilibrium systems 213Bi + 221Fr and 213Bi + 225Ac were prepared. For the measurement conditions used, the equilibrium spectra of 225Ac with daughter decay products were found to be unsuitable for accurate determination of the energy of the 213Bi major peak and can be used only for tentative estimations. The actual energies of 213Bi α-radiation do not coincide with the assessed reference data used today and are in the region of the upper limit of the uncertainty range for these data. The results show that the presently used energy characteristics of 213Bi α-radiation require refinement.  相似文献   

5.
Distribution of 210Po in thalli of soil and wood (epiphytic) lichens was studied. Four fractions containing the corresponding 210Ро forms were obtained by sequential extraction: (1) intercellular, (2) extracellular, (3) intracellular, and (4) insoluble thallus residue. The 210Ро uptake by lichens is mainly passive, as the total content of the radionuclide in fractions 1, 2, and 4 reaches 88–97%. From 3 to 12% of 210Ро is taken up actively (fraction 3), and for soil lichens this parameter is approximately 2.75 times higher than for epiphytic lichens. Presumably, 210Ро is supplied into soil and epiphytic lichens in the form of different chemical compounds and is therefore characterized by different bioavailability.  相似文献   

6.
In this study, ZrP2O7 was synthesized by the solid state reaction of ZrO2 and NH4H2PO4 at 900 °C. Then, in set 1; 10, 5, 1, 0.5, 0.1, 0.05, 0.03% previously prepared Sr2P2O7 were doped into ZrP2O7, and Sr2P2O7 slightly affect the unit cell parameter of cubic ZrP2O7 (a = 8.248(6)–8.233(8) Å). The reverse of this process was also applied to Sr2P2O7 system (set 2). ZrP2O7 changes the unit cell parameters of orthorhombic Sr2P2O7 in between a = 8.909(5)–8.877(5) Å, b = 13.163(3)–13.12(1) Å, and c = 5.403(2)–5.386(4) Å. Analysis of the vibrations of the P2O 7 4? ion and approximate band assignments for IR and Raman spectra are also reported in this work. Some coincidences in infrared and Raman spectra both sets were found and strong P–O–P bands were observed. Surface morphology, EDX analysis, and thermoluminescence properties of both sets were given the first time in this paper.  相似文献   

7.
This paper reports a study on the mechanical and tribological properties of ab- and a (b) c?planes of YBa2Cu3O7?δ single crystals. The single crystals were grown using a CuO-BaO self-flux method. The oxygenation effect on the mechanical and tribological properties of ab- and a (b) c?planes is reported. For the ab- plane, the hardness and elastic modulus were around 6 and 50 GPa, respectively. In this case, significant differences were not observed among the hardness and elastic modulus at different oxygenation states. However, the hardness and elastic modulus for as-grown and oxygenated YBa2Cu3O7?δ single crystals were different from that of the a (b) c?plane, and were observed to be slightly higher for the as-grown than for the oxygenated samples. For as-grown and oxygenated samples, we observed hardness values around 4.7 and 2.0 GPa, respectively. Regarding the elastic modulus, the values were 75 and 40 GPa, respectively. The indentation fracture toughness values on the ab- plane for the as-grown and oxygenated YBa2Cu3O7?δ single crystal were 3.7 ± 1.2 and 2.9 ± 1.2 MPa m1/2, respectively. For the ab- plane, the scratch resistance of the as-grown sample was higher than that of the oxygenated sample and the scratches under load were deeper for the oxygenated sample. As regards the a (b) c?plane, the scar features were seemingly constant through all the scratch lengths and the scratches under load were deeper and larger for the oxygenated than that for the as-grown sample.  相似文献   

8.
The kinetics of thermal dehydration of Mg3(PO4)2 · 8H2O was investigated using thermogravimetry at four different heating rates. The activation energies of the dehydration step of Mg3(PO4)2 · 8H2O were calculated through the isoconversional Ozawa and Kissinger-Akahira-Sunose (KAS) methods and iterative methods, which were found to be consistent and indicate a single mechanism. The possible conversion function of the dehydration reaction for Mg3(PO4)2 · 8H2O has been estimated through the Coats and Redfern integral equation, and a better kinetic model such as random nucleation of the “Avrami–Erofeev equation (A 3/2 model)” was found. The thermodynamic functions (ΔH*, ΔG*, and ΔS*) of the dehydration reaction are calculated by the activated complex theory and indicate that it is a non-spontaneous process when the introduction of heat is not connected.  相似文献   

9.
We have synthesized nanoparticulate cobalt(II) hydroxide containing Co2+ in tetrahedral oxygen coordination (Co Td 2+ ), atypical of such systems: nano- [Co(OH)2(H3O) δ + ]δ+. The (Co Td 2+ ) coordination in the hydroxide is inferred from its electronic diffuse reflectance spectrum, which shows a multiplet of strong absorption bands at 14500, 15000, and 16000 cm?1 (4 A 2(F)-4 T 1(P) transition). Nanoparticulate cobalt(II) hydroxide forms in a weakly acidic medium under essentially nonequilibrium conditions due to supersaturation (by three to four orders of magnitude) with the starting reagents (CoCl2 and LiOH) at the instant of the formation of the poorly soluble phase Co(OH)2. Presumably, colloidal particles of nanoparticulate cobalt(II) hydroxide in a weakly acidic aqueous medium have a positive surface charge, compensated by a counter-ion (Cl?) layer: nano-[Co(OH)2(H3O) δ + ]δ+ · δCl?. The XRD patterns of pastes (gels) containing this hydroxide show three broad-ened lines with d = 5.31 (2θ = 16.7°), 2.77 (2θ = 32.3°), and 2.32 Å (2θ = 38.8°). According to small-angle X-ray scattering data, nano-[Co(OH)2(H3O) δ + ]δ+ has a narrow particle size distribution (1.0–2.0 nm). Synthesis and storage conditions are identified which ensure stabilization of the electronic state and particle size of nano-[Co(OH)2(H3O) δ + ]δ+ for a long time.  相似文献   

10.
The kinetics of spontaneous demagnetization in nanoparticles of the exotic epsilon-phase of indium-doped iron(III) oxide (ε-In0.24Fe1.76O3) has been studied using the method of accelerated testing of magnets for temporal stability in a magnetization-reversal field. Time dependences of the magnetization of nanoparticles measured in a wide range of magnetic fields exhibited rectification in semilogarithmic coordinates. The dependence of the magnetic viscosity on the magnetic field has been measured and used for determining the fluctuation field and activation volume. A relationship between the magnetic viscosity and magnetic noise caused by random thermoinduced magnetization reversal in separate nanoparticles is established.  相似文献   

11.
237U was produced by the reaction 238U(γ, n) on an electron accelerator, MT-25 microtron, at the Flerov Laboratory of Nuclear Reactions. For the separation of 237U and [238U, the recoil nuclei were collected by a nanostructured material, hydrated manganese dioxide (of the cryptomelane type), in the solid-solid system. From fission products, 237U was separated by ion exchange. The specific activity of the resulting 237U was 4.5 × 109 Bq (mg 238U)-1, with the content of radioactive impurities of ≤10-6 Bq Bq-1. The chemical yield of 237U was 80%.  相似文献   

12.
In this study, we confirmed that the characteristics of anion intercalation into the interlayer of a hydrotalcite-like compound (HT) during synthesis are similar to those of the anion-exchange reaction of HTs as well as the reconstruction reaction of HTs from Mg-Al oxide. We demonstrated that (i) Cl, which has a higher charge density than NO3, more easily reacted with Mg and Al species to form HT structure, resulting in greater intercalation of Cl into the HT interlayer; and (ii) for HTs with lower Mg: Al molar ratios, OH, which has a higher charge density than Cl and NO3, was more likely to interact with Mg and Al species to form HT structure, blocking the intercalation of Cl and NO3. Furthermore, we showed that high concentrations of Cl and NO3 in solution regulated their intercalation into the HT interlayer. The high activity of Cl and NO3 in solution would facilitate the anions’ reactions with Mg and Al species to form HTs, resulting in a high degree of anion intercalation into the interlayer of HTs.  相似文献   

13.
This paper presents results of a 57Fe probe Mössbauer spectroscopy study of the BiNi0.9657Fe0.04O3 nickelate. The spectra measured above its TN demonstrate that Fe3+ cations heterovalently substitute for Ni2+ nickel (←Fe3+), being stabilized on four sites of the nickel sublattice in the structure of BiNiO3. Calculations in an ionic model with allowance for monopole and dipole contributions to the electric field gradient indicate that the parameters of electric hyperfine interactions between 57Fe probe atom nuclei reflect the specifics of the local environment of the nickel in the structure of the unsubstituted BiNiO3 nickelate. Below TN, Mössbauer spectra transform into a complex Zeeman structure, which is analyzed in terms of first-order perturbation theory with allowance for electric quadrupole interactions as a small perturbation of the Zeeman levels of the 57Fe hyperfine structure, as well as for specific features of the magnetic ordering of the Ni2+ cations in the nickelate studied.  相似文献   

14.
We have studied in detail the gamma radiation induced changes in the electrical properties of the (TeO2)0·9 (In2O3)0·1 thin films of different thicknesses, prepared by thermal evaporation in vacuum. The current–voltage characteristics for the as-deposited and exposed thin films were analysed to obtain current versus dose plots at different applied voltages. These plots clearly show that the current increases quite linearly with the radiation dose over a wide range and that the range of doses is higher for the thicker films. Beyond certain dose (a quantity dependent on the film thickness), however, the current has been observed to decrease. In order to understand the dose dependence of the current, we analysed the optical absorption spectra for the as-deposited and exposed thin films to obtain the dose dependences of the optical bandgap and energy width of band tails of the localized states. The increase of the current with the gamma radiation dose may be attributed partly to the healing effect and partly to the lowering of the optical bandgap. Attempts are on to understand the decrease in the current at higher doses. Employing dose dependence of the current, some real-time gamma radiation dosimeters have been prepared, which have been found to possess sensitivity in the range 5–55 μGy/μA/cm2. These values are far superior to any presently available real-time gamma radiation dosimeter.  相似文献   

15.
The effect of experimental conditions on UVL2 (1–2 mM) disproportionation was studied spectrophotometrically through the UIVL2 accumulation (L = P2W17O 61 10? ). In 1 M NaNO3 solution containing 0.01 M HAc and 0.01 M NaAc, the rate of UVL2 disappearance is described by the equation V = k 1[UVL2]. The k 1 value is almost constant with pH decreasing from 4.5 to 1.7, but increases with increasing acetate concentration; the presence of 1 mM UIVL2, U(VI), or L does not affect k 1. In the solutions of 0.1–1.0 M HClO4 (ionic strength 1), the reaction rate is described by the equation V = 2k 2[H+]2.5[UVL2]2. Probable disproportionation mechanism is discussed. The first stage is substitution of L by water molecules in the UIVL2 complex and appearance of the reactive U(V) complex with mixed coordination sphere.  相似文献   

16.
The potential of the Fe(CN)63−/Fe(CN)64− couple in solutions containing 0.5–1 M alkali and high concentrations of Li, Na, and K salts was estimated potentiometrically. The reactions of Fe(CN)63− with Np(VI) in such media were studied by spectrophotometry. In 0.5–1 M KOH + 7 M KF or 0.5–1 M KOH + 8 solutions, Np(VI) is oxidized to the heptavalent state with a small excess of Fe(CN)63−. In the presence of lithium or sodium salts, a larger excess of the oxidant is required for complete oxidation of Np(VI). Carbonate ions in concentrations exceeding 2 M prevent the Np(VI) oxidation.  相似文献   

17.
A specific feature of U(IV) oxidation with xenon difluoride in aqueous H2SO4 solutions is low-temperature chemiluminescence (CL), which in the course of warming the sample quickly cooled to 77 K is recorded starting from 165 K and reaches a maximum at about 200 K. Exothermic phase transitions, crystallization of the ice + H2SO4·4H2O and ice + H2SO4·6.5H2O eutectics, occur in the same temperature range. The data (temperature dependences of the chemiluminescence intensity and simultaneously recorded DTA curves) obtained in experiments with variation of the rate of mixing and cooling the solutions and of the concentrations of H2SO4 and F? and UO 2 2+ ions are well explicable by the catalytic activity of the juvenile surface of H2SO4 crystal hydrates toward low-temperature reaction of U(IV) with XeF2.  相似文献   

18.
The effect of microwave radiation (MWR) on the decomposition of UO2(NO3)2·6H2O was studied. Determination of [UO 2 2+ ] and [NO 3 ? ], and also of the molar ratio NO 3 ? : UO 2 2+ in various fractions of the decomposition product showed that the mechanism of the UO2(NO3)2·6H2O decomposition under the action of MWR differs from the mechanism of its decomposition under common convection heating. The main precursor of UO3 as product of UO2(NO3)2·6H2O decomposition under the action of MWR is uranyl hydroxonitrate UO2(OH)NO3 formed already in the first minutes of the irradiation. In contrast to the thermolysis under convection heating, UO2(NO3)2 or its hydrates were not detected as intermediates. The mechanism of the UO2(NO3)2·6H2O decomposition under the action of MWR can be presented by the reactions UO2(NO3)2·6H2O → UO2(OH)NO3 + 5H2O + HNO3 and UO2(OH)NO3 → UO3 + HNO3. The solubility of UO2(OH)NO3 in H2O at 20°C was estimated experimentally at 6.83 × 10?2 M.  相似文献   

19.
We have investigated the interaction between (Bi,Pb)2Sr2Ca2Cu3O10+δ (Bi-2223) and small additions (0.05–0.3 wt %) of nitride powders (TaN, AlN, HfN, NbN, Si3N4, TiN, and ZrN) with a particle size from 0.02 to above 0.5 μm and the effect of these nitrides on the microstructure, phase composition, distribution, and morphology of the resulting second-phase inclusions. The concentration and particle size of the nitrides and sintering conditions are shown to influence the superconducting transition temperature T c, critical current density j c, irreversible remanent magnetization, bulk density, and mechanical properties of the Bi-2223/nitride composites.  相似文献   

20.
In this study, we report the results of an investigation into the sintering temperature dependence of magnetic and transport properties for GdBaCo2 O 5 + δ synthesized through a sol-gel method. The lowering of sintering temperature leads to the increase of oxygen content and the reduction of grain size. The increase of oxygen content results in the enhancement of magnetic interactions and the weakening of Coulomb repulsion effect, while the reduction of grain size improves the magnetoresistance effect. Metal-insulator transition accompanied with spin-state transition is observed in all samples.  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