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.
A. Billard  P. Vernoux 《Ionics》2005,11(1-2):126-131
8 to 120 nm-thick Pt coatings were sputter-deposited on Yttria-Stabilised Zirconia (YSZ) membranes, 17 mm in diameter, by magnetron sputtering of a Pt target at low pressure (0.3 Pa). The catalytic activity of propane combustion under open-circuit conditions is first measured near to the stoichiometry (0.2%C3H8/1%O2) and shows that the coatings present a rather high catalytic activity. Close-circuit measurements were finally performed at 337 °C and 400 °C. They show that rather high faradaic efficiencies in the range 106 can be reached as soon as the Pt film is thick enough to allow a bias of its whole area. Paper presented at the Patras Conference on Solid State Ionics — Transport Properties, Patras, Greece, Sept. 14 – 18, 2004.  相似文献   

2.
The sonocatalytic degradation of EDTA (C0 = 5 10−3 M) in aqueous solutions was studied under 345 kHz (Pac = 0.25 W mL−1) ultrasound at 22–51 °C, Ar/20%O2, Ar or air, and in the presence of metallic titanium (Ti0) or core-shell Ti@TiO2 nanoparticles (NPs). Ti@TiO2 NPs have been obtained using simultaneous action of hydrothermal conditions (100–214 °C, autogenic pressure P = 1.0–19.0 bar) and 20 kHz ultrasound, called sonohydrothermal (SHT) treatment, on Ti0 NPs in pure water. Ti0 is composed of quasi-spherical particles (30–150 nm) of metallic titanium coated with a metastable titanium suboxide Ti3O. SHT treatment at 150–214 °C leads to the oxidation of Ti3O and partial oxidation of Ti0 and formation of nanocrystalline shell (10–20 nm) composed of TiO2 anatase. It was found that Ti0 NPs do not exhibit catalytic activity in the absence of ultrasound. Moreover, Ti0 NPs remain inactive under ultrasound in the absence of oxygen. However, significant acceleration of EDTA degradation was achieved during sonication in the presence of Ti0 NPs and Ar/20%O2 gas mixture. Coating of Ti0 with TiO2 nanocrystalline shell reduces sonocatalytic activity. Pristine TiO2 anatase nanoparticles do not show a sonocatalytic activity in studied system. Suggested mechanism of EDTA sonocatalytic degradation involves two reaction pathways: (i) sonochemical oxidation of EDTA by OH/HO2 radicals in solution and (ii) EDTA oxidation at the surface of Ti0 NPs in the presence of oxygen activated by cavitation event. Ultrasonic activation most probably occurs due to the local heating of Ti0/O2 species at cavitation bubble/solution interface.  相似文献   

3.
Absorption cross sections of molecular oxygen were measured at the H and D Ly-α wavelengths (1215.67 and 1215.34 Å) between 800 and 1700°K, the cross sections being given by the equation σ (cm2) = 4.2E-18 exp(-3070/T) (4.2E-18 = 4.2 × 10-18). The absorption cross section for v = 1 is (9±2)E-19 cm2 and that for v = 2 is (7±3)E-18 cm2, compared to 1.E-20 cm2 for v = 0. Vibrational relaxation times were found to be in agreement with literature data over the range of common temperatures.  相似文献   

4.
In this paper, ZnO microspheres, which are composed of irregular nanoparticles, have been synthesized successfully from a metal-organic precursor. The average diameter is about 3.5 μm and the specific surface area is 7.53 m2 g?1. Measured by electrochemical tests as electrode materials for supercapacitors, the ZnO powders show high specific capacitances (1017.5 Fg?1 at 5 Ag?1 and 562.5 Fg?1 at 50 Ag?1, respectively) and excellent cycling stability (the specific capacitance was kept at 631.2 Fg?1 and 89.2 % retention after 3000 cycles at 18 Ag?1). These results show that the microspherical ZnO could be a potential electrode material for supercapacitors.  相似文献   

5.
The 18N(β?)18O decay was observed via the 9Be(18O, p2α)18N reaction, utilizing helium-transport techniques and Ge(Li) spectroscopy. In addition to the previously reported β-decay to the 18O 4456 keV level (Jπ = 1?) branches were observed to levels at excitation energies (in keV) of 1982 (Jπ = 2+). 5530(2?), 6198(1?). 6350(2? or 1+), 6880(0?), and 7771(2?). The percentage β-branches, in order of increasing excitation energy, are 3.9 ± 1.5, 54.6 ± 1.0, 3.1 ± 0.4, 1.4 ± 0.2, 2.2 ± 0.3, 14.8 ± 0.8 and 5.0 ± 0.5. respectively, with 15 % assumed on the basis of calculations to proceed to non-γ-emitting states. These measurements allow definite assignments Jπ = 1? for the 18N ground state and Jπ = 0? for the 18O 6880 keV state. Additional measurements determine the 18N half-life to be T1 2 = 624 ± 12 ms. A shell-model calculation for mass 18 was carried out in a full 1?ω basis. The predictions for the T = 2 energy level spectrum and for 18N β-decay are discussed.  相似文献   

6.
Abstract

In acetonitrile solutions, the exchange reaction is bimolecular in the Tl+ + 18C6 system, while in the Tl+ + pentaglyme system the associative-dissociative and the bimolecular mechanisms coexist at room temperature and the bimolecular exchange reaction dominates at 263° K. For the bimolecular mechanism in the case of Tl+ + 18C6 and the associative-dissociative mechanism in the case of Tl+ + pentaglyme, the activation energies of the exchange reactions change with temperature. At 298° K, in the Tl+ + 18C6 system the activation energy for the bimolecular exchange reaction is ≈ 2 kcal.mol?1 and exchange rate constant (k1) is (4.1 ± 0.1) × 107 s?1mol?1; in the Tl+ + pentaglyme system, the activation energy for the associative-dissociative exchange reaction is ≈ 5 kcal mol?1 and the decomplexation rate constant (k?2) is (2.2 ± 0.4) X 105 s?1. The activation energy for the bimolecular exchange in the Tl+ + pentaglyme system was determined to be 3.00 ± 0.05 kcal.mol?1 and the exchange rate constant (3.0 ± 0.1) X 108 s?1 mol?1.  相似文献   

7.
The mass of 18C has been measured using the double-charge-exchange reaction 48Ca(18O, 18C)48Ti at an 18O energy of 112 MeV. The 18C ions were detected at the focal plane of a magnetic spectrometer. The mass excess of 18C was found to be 24.923 ± 0.030 MeV, and the first excited state was observed at an excitation energy of 1.62 ± 0.02 MeV. At the same time, an independent measurement of the mass excess of 17C was obtained from the 48Ca(18O, 17C)49Ti reaction, and the value 21.039 ± 0.020 MeV is in excellent agreement with an earlier measurement. The first excited state of 17C is at 295 ± 10 keV.  相似文献   

8.
Cross sections have been measured for emission of the 2nd positive system of N2 and the 1st negative system of N+2 by electron excitation. The electron energy ranged from the threshold to 400 eV. To obtain absolute values, comparison with the emission cross section of the 51S He level has been used.The results for the (0, 2) band of the 2nd positive system and for the (0, 1) band of the 1st negative system are (2.87 ± 0.40) × 10-18 and (5.35 ± 0.5) × 10-18cm2, respectively, at the peak. The cross sections are given for 23 bands of the 1st negative systems.Use of the (0, 1) band of the 1st negative system is proposed for reference in obtaining absolute values.  相似文献   

9.
ABSTRACT

Karst springs in the Main Range of the Crimean Mountains and the Crimean Piedmont show a restricted range of values (δ18O?=?–10.5 to –8.0 ‰, δ2H?=?–72 to –58 ‰), somewhat more negative than the weighted mean of meteoric precipitation. This suggests preferential recharge at higher elevations during winter months. Groundwater tapped by boreholes splits in three groups. A first group has isotopic properties similar to those of the springs. The second group shows significantly lower values (δ18O?=?–13.3 to –12.0 ‰, δ2H?=?–95 to –82 ‰), suggesting recharge during colder Pleistocene times. The third group has high isotope values (δ18O?=?–2.5 to +1.0 ‰, δ2H?=?–24 to –22 ‰); the data points are shifted to the right of the Local Meteoric Water Line, suggesting water–rock exchange processes in the aquifer. These boreholes are located in the Crimean Plains and discharge mineralized (ca. 25 g L?1) thermal (65°C) water from a depth of 1600–1800 m. Groundwater associated with mud volcanoes on the Kerch peninsula have distinct isotope characteristics (δ18O?=?–1.6 to +9.4 ‰, δ2H?=?–30 to –18 ‰). Restricted δ2H variability along with variable and high δ18O values suggest water–rock interactions at temperatures exceeding 95 °C.  相似文献   

10.
Abstract

The defects in n-GaP crystals irradiated by 2.3 MeV electrons up to 1 × 1019 cm?2 at RT were studied by means of positron annihilation (angular correlation) and electrical property measurements. It was found that positrons are trapped in some radiation-induced vacancy-type defects (acceptors) but that the effect saturates at high electron fluences (D1 × 1018 cm?2). The trapping rate in irradiated samples increases with temperature in the range 77–300 K. Post-irradiation isochronal annealing reveals the positron traps clustering at about 200–280°C. All positron sensitive radiation-induced defects disappear upon annealing up to 500°C.  相似文献   

11.
ABSTRACT

Tungsten (W) has been regarded as one of the most promising plasma facing materials (PFMs) in fusion reactors. The formation of bubbles and blisters during hydrogen (H) irradiation will affect the properties of W. The dependence of implantation conditions, such as fluence and energy, is therefore of great interest. In this work, polycrystalline tungsten samples were separated into two groups for study. The thick samples were implanted by 18?keV H3+ ions to fluences of 1?×?1018, 1?×?1019 and 1?×?1020 H+/cm2, respectively. Another thick sample was also implanted by 80?keV H2+ ions to a fluence of 2?×?1017 H+/cm2 for comparison. Moreover, the thin samples were implanted by 18?keV H3+ ions to fluences of 9.38?×?1016, 1.88?×?1017 and 5.63?×?1017 H+/cm2, respectively. Focused ion beam (FIB) combined with scanning electron microscopy (SEM) and transmission electron microscopy (TEM) were used for micro-structure analysis, while time-of-flight ion mass spectrometry (ToF-SIMS) was used to characterize the H depth profile. It is indicated that bubbles and blisters could form successively with increasing H+ fluence. H bubbles are formed at a fluence of ~5.63?×?1017 H+/cm2, and H blisters are formed at ~1?×?1019 H+/cm2 for 18?keV H3+ implantation. On the other hand, 80?keV H2+ ions can create more trapping sites in a shallow projected range, and thus enhancing the blisters formation with a relatively lower fluence of 2?×?1017?H+/cm2. The crack-like microstructures beneath the blisters are also observed and prefer to form on the deep side of the implanted range.  相似文献   

12.
The neutron irradiated reactor pressure vessel (RPV) steels at various dose of 0–1018 n/cm2 have been studied with Mössbauer, x-ray diffraction, and VSM. The Mössbauer data shows that the value of magnetic hyperfine field of Fe atom that exist at martensite is 330 kOe at site 1 and 305 kOe at site 2. At room temperature, the total absorption area of Mössbauer spectra with respect to irradiation of neutron is constant for the dose of 0–1016 n/cm2, while over the dose of 0–1017 n/cm2 the absorption area decreases rapidly. But the doublet area for the dose of 0–1016 n/cm2 is constant, while over the dose of 1017 n/cm2 it increases with increasing the fluence level of neutron. The coercivity and remanence of the neutron irradiated samples do not change significantly. But the maximum induction decreases by 5% at 1018 n/cm2, compared with that of the as-received sample.  相似文献   

13.
Oxygen isotope separation has been examined by utilizing the two-frequency infrared multiphoton dissociation (IRMPD) of 2,3-dihydropyran (DHP). The two-frequency IRMPD reduces the required laser fluences to those lower than the damage threshold of optical windows. For example, dissociation probability of DHP containing 18O (D(18O)) and enrichment factor (S) were obtained to be 1.6×10−3/pulse and 316, respectively, by the simultaneous irradiation with 1052.2 cm−1 photons at 0.45 J/cm2 and 1031.5 cm−1 photons at 1.06 J/cm2. These are comparable with D(18O)=2.2×10−3/pulse and S=391 obtained by the single-frequency irradiation of 1033.5 cm−1 photons at 2.2 J/cm2. Therefore, the production rate of an 18O enriched dissociation product has been increased to four times or more, compared with the single-frequency IRMPD, and this two-frequency method would promise a practical large scale separation.  相似文献   

14.
In this paper, we report investigations of the electron emission from pure Zn cathodes irradiated by UV laser pulses of 23 ns (full-width at half-maximum) at a wavelength of 248 nm (5 eV). The metal cathodes were tested in a vacuum photodiode chamber at 10?5 Pa. They were irradiated at normal incidence and the anode–cathode distance was set at 3 mm. The maximum applied accelerating voltage was 18 kV, limited by the electrical breakdown of the photodiode gap. Under the above experimental conditions, a maximum applied electric field of 6 MV/m resulted. In the saturation regime, the measured quantum efficiency value increased with the accelerating voltage due to the plasma formation. The highest output current was achieved with 14 mJ laser energy, 18 kV accelerating voltage and its value was 12 A, corresponding to a global quantum efficiency (GQE) approximately of 1×10?4. The temporal quantum efficiency was 1.0×10?4 at the laser pulse onset time and 1.4×10?4 at the pulse tail. We calculated the target temperature at the maximum laser energy. Its value allowed us to obtain output pulses of the same laser temporal profile. Tests performed with a lower laser photon energy (4.02 eV) demonstrated a GQE of two orders of magnitude lower.  相似文献   

15.
ABSTRACT

A highly sensitive, selective, and rapid method for the determination of ng mL?1 level of U(VI) based on the rapid reaction of U(VI) with 2-(2-benzothiazolylazo)-3-hydroxyphenol (BTAHP) and the solid-phase extraction of the colored complex with a reversed-phase polymer-based C18 cartridge was developed. The BTAHP reacted with U(VI) to form a violet complex of molar ratio 2:1 [BTAHP to U(VI)] in the presence of 4.0 M of phosphoric acid solution and Triton X-114 medium. This complex was enriched by the solid-phase extraction with a polymer-based C18 cartridge. The enrichment factor of 200 was achieved. The molar absorptivity of the complex is 2.73 × 106 L mol?1 cm?1 at 639 nm in the measured solution. The system obeys Beer's law in the range of 2.0–125 ng mL?1, whereas the optimum concentration range obtained from Ringbom plot was 8.0–115 ng mL?1. The relative standard deviation for 10-replicates sample of 100 ng mL?1 level is 1.05%. The detection and quantification limits are 0.6 and 1.98 ng mL?1 in the original sample. This method was applied to the determination of uranium(VI) in sea, tap, and waste waters, ore standard reference material, soil and sediment samples with good results comparing to the graphite furnace atomic absorption spectroscopy (GFAAS) method.  相似文献   

16.
Using the time-differential perturbed-angular correlation technique, we have observed the nuclear electric quadrupole interaction at 181Ta (482 keV) as an impurity in the refractory compound HfB2. The measured interaction frequency is νQ=730±5 MHz which corresponds to an electric field gradient of |eq|=(1·19±0·05)×1018 V/cm2 at room temperature. By considering a second measurement at 4·2°K, which yields the same results as above, and by comparison with available results for Hf in HfB2, we conclude that the d-electron density of states at the Fermi level is quite small in agreement with trends observed by others.  相似文献   

17.
The potential for using Raman spectroscopy to measure stable oxygen isotope ratios (18O/16O) in carbonates is evaluated by measuring the Raman spectra and isotope ratios of a suite of 60 synthesized, 18O‐enriched calcite crystals ranging in composition from natural abundance (0.2 mole‐% 18O) to 1.2 mole‐% 18O. We determined the Raman‐inferred isotopic ratios (RRaman) by fitting curves to the ν1 symmetric stretching peak at 1086 cm−1 and the smaller satellite peak, associated with the ν1 stretching mode of singly substituted carbonate groups (C16O218O) at 1065 cm−1. The ratio of the two peak areas shows a 1:1 correspondence with the 18O/16O ratios derived from standard mass spectrometry methods, confirming that the relative intensities of the ν1 symmetric stretching peaks is a direct measure of the isotopic ratio in the carbonates. The 1‐sigma uncertainties of the RRaman values of the individual crystals were 0.00079 (384‰ PDB) and 0.00043 (210‰ PDB) for the four‐crystal sample means. This level of uncertainty is much too high to provide significant estimates of natural variability; however, there are multiple prospects for improving the accuracy and precision of the technique. Carbon isotope ratios in carbonates cannot be measured by our approach, but our results highlight the potential of Raman‐based isotope ratio measurement for C and other elements in minerals and organic compounds. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

18.
Two types of commercially available carbon fibers (high tensile strength, HTS, and high modulus, HMS) were irradiated in the Ground Test Reactor in environments of air and of liquid nitrogen (LN2). The tensile strength of HTS fibers irradiated in air increased sharply above a fast-neutron fluence of 6 × 1017 n/cm2 (E> 1 MeV) and was 17 per cent greater than the strength of unirradiated control fibers at a fluence of 8.5 × 1017 n/cm2, but then the strengrh began to decrease for additional neutron exposure in air and fell 25 per cent below the control strength at the highest fluence of 4.5 × 1018 n/cm2. However, when irradiated in LN2 where surface oxidation did not take place, the room-temperature strength of HTS fibers continued to increase beyond 8.5 × 1017 n/cm2 and became almost 30 per cent greater than the control strength for a fluence of 3 × 1018 n/cm2. The tensile strength of HMS fibers irradiated in air increased slowly but steadily with neutron exposure and was only 4 per cent greater than the control strength at the highest fluence of 4.5 × 1018 n/cm2; the room-temperature strength of the HMS fibers decreased by 13 per cent when irradiated to a fluence of 3 × 1018 n/cm2 in LN2.  相似文献   

19.
Comparison of the phonons and electronic excitations of UO2 with different oxygen isotopes (16O, 18O), two purely electronic excitations at 4162 and 5778 cm−1 have been evidenced by their limited shift, when 16O is replaced by 18O. The bands observed at 570, 1144 and 2300 cm−1 in U16O2 shift to 545, 1091 and 2190 cm−1 for U18O2. The wavenumbers and the isotope shifts of these bands exhibit multiple relationships. It supports the assignments of the 1LO (~570 cm−1) and 2LO (~1144 cm−1) phonons for U16O2. Both bands exhibit marked decrease for increasing O/U ratio. The observed phenomenon is regarded as a good estimation of the stoichiometry of a uranium oxide. In addition, by analyzing the intensity ratios of the 2LO to 1LO phonons for various UO2 samples, the 2LO band is assigned to the Fröhlich interaction. Defect‐induced and/or deformation potential mechanisms play a dominant role for the 1LO band. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

20.
Abstract

Niobium samples were neutron-irradiated at reactor ambient temperatures (approximately 50 °C for polycrystals and 90°C for single crystals) to doses from 4 × 1015 to 8 × 1018 neutrons/cm2 (E> 1 MeV). The density and size distribution of radiation-produced defect clusters, observed by transmission electron microscopy, were measured in polycrystalline niobium over a range of doses from 2 × 1017 to 4.4 × 10l8 neutrons/cm2. The TEM results were correlated with yield stress measurements as a function of dose. The radiation hardening was analyzed on the basis of a planar dispersed barrier model. It was found that the observed increases in yield stress at low doses were consistent with the measured density and size distribution of the defect clusters, with a cluster strength of (0.5–0.8) Gb2 (G, shear modulus; b, Burgers vector). This corresponds to strong barrier hardening. At doses above about 1018 neutrons/cm2, the hardening rate decreased sharply; this apparent saturation is discussed in terms of the coarsening of defect clusters, dislocation channeling, and the effect of interstitial impurities.  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