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.
Single crystals of z- and x-cut LiNbO3 were irradiated at room temperature and 15 K using He+- and Ar+-ions with energies of 40 and 350 keV and ion fluences between 5 × 1012 and 5 × 1016 cm−2. The damage formation investigated with Rutherford backscattering spectrometry (RBS) channeling analysis depends on the irradiation temperature as well as the ion species. For instance, He+-irradiation of z-cut material at 300 K provokes complete amorphization at 2.0 dpa (displacements per target atom). In contrast, 0.4 dpa is sufficient to amorphize the LiNbO3 in the case of Ar+-irradiation. Irradiation at 15 K reduces the number of displacements per atom necessary for amorphization. To study the etching behavior, 400 nm thick amorphous layers were generated via multiple irradiation with He+- and Ar+-ions of different energies and fluences. Etching was performed in a 3.6% hydrofluoric (HF) solution at 40 °C. Although the etching rate of the perfect crystal is negligible, that of the amorphized regions amounts to 80 nm min−1. The influence of the ion species, the fluence, the irradiation temperature and subsequent thermal treatment on damage and etching of LiNbO3 are discussed.  相似文献   

2.
We have used positron Doppler-broadening spectroscopy to examine a series of neutron-irradiated model alloys (1 × 1023 n/m2, E>0.5 MeV) and 73W-weld steel (to 1.8 × 1023 n/m2, E>1 MeV. The copper, nickel and phosphorus content of the model alloys was systematically varied. The samples were examined in the as-irradiated state and after post-irradiation isochronal anneals to temperature up to 600 °C. By following the S and W parameters, and especially by plotting the results in (S,W) space, we can infer that the damage is a combination of irradiation-induced metallic precipitates and vacancy-type defect clusters. Samples with either high Cu or with a combination of high Ni and medium Cu (and the pressure-vessel weld steel) showed evidence for both irradiation-induced metallic precipitation, and vacancy-type clusters, while samples without either high Ni or high Cu showed predominantly evidence of annihilations in vacancy-type clusters. These results are discussed in terms of embrittlement models.  相似文献   

3.
Ni+ ion implantation with an energy of 64 keV in MgO single crystals was conducted at room temperature up to a fluence of 1 × 1017 ion/cm2. The as-implanted crystals were annealed isochronally at temperatures up to 900 °C. Optical absorption spectroscopy, X-ray photoelectron spectroscopy (XPS) and transmission electron microscopy (TEM) have been utilized to characterize the changes of optical properties and the microstructure of the annealed samples. XPS results showed that the charge state of implanted Ni was still mainly in metallic Ni0 after annealing at 900 °C. TEM analysis revealed metallic Ni nanoparticles with depth-dependant dimensions of 1–10 nm in the annealed sample. Optical absorption spectroscopy indicated that the Ni nanoparticles exhibited a broad surface plasmon resonance absorption band in annealed samples and the band shifted to a longer wavelength with the increasing annealing temperature.  相似文献   

4.
The temperature dependence of ion-induced electron emission yield γ under 30 keV Ar+ ion impacts at incidence angles θ = 0−80° under dynamically steady-state conditions has been measured for polygranular graphite POCO-AXF-5Q. The fluencies were 1018–1019 ion/cm2, the temperatures varied from the room temperature (RT) to 400 °C. The RHEED has shown that same diffraction patterns correspond to a high degree of disorder at RT. At high temperature (HT), some patterns have been found similar to those for the initial graphite surfaces. The dependence γ(T) has been found to be non-monotonic and for normal and near normal ion incidence manifests a step-like increase typical for a radiation induced phase transition. At oblique and grazing incidence (θ > 30°), a broad peak was found at Tp = 100 °C. An analysis based on the theory of kinetic ion-induced electron emission connects the behavior of γ(θ,T) to the dependence of both secondary electron path length λ and primary ion ionizing path length Re on lattice structure that drastically changes due to damage annealing.  相似文献   

5.
Xe+ ion implantation with 200 keV was completed at room temperature up to a fluence of 1 × 1017 ion/cm2 in yttria-stabilized zirconia (YSZ) single crystals. Optical absorption and X-ray photoelectron spectroscopy (XPS) were used to characterize the changes of optical properties and charge state in the as-implanted and annealed crystals. A broad absorption band centered at 522 or 497 nm was observed in the optical absorption spectra of samples implanted with fluences of 1 × 1016 ion/cm2 and 1 × 1017 ion/cm2, respectively. These two absorption bands both disappeared due to recombination of color centers after annealing at 250 °C. XPS measurements showed two Gaussian components of O1s spectrum assigned to Zr–O and Y–O, respectively, in YSZ single crystals. After ion implantation, these two peaks merged into a single peak with the increasing etching depth. However, this single peak split into two Gaussian components again after annealing at 250 °C. The concentration of Xe decreased drastically after annealing at 900 °C. And the XPS measurement barely detected the Xe. There was no change in the photoluminescence of YSZ single crystals with a fluence of 1 × 1017 ion/cm2 after annealing up to 900 °C.  相似文献   

6.
In the present study, a 500 Å thin Ag film was deposited by thermal evaporation on 5% HF etched Si(1 1 1) substrate at a chamber pressure of 8×10−6 mbar. The films were irradiated with 100 keV Ar+ ions at room temperature (RT) and at elevated temperatures to a fluence of 1×1016 cm−2 at a flux of 5.55×1012 ions/cm2/s. Surface morphology of the Ar ion-irradiated Ag/Si(1 1 1) system was investigated using scanning electron microscopy (SEM). A percolation network pattern was observed when the film was irradiated at 200°C and 400°C. The fractal dimension of the percolated pattern was higher in the sample irradiated at 400°C compared to the one irradiated at 200°C. The percolation network is still observed in the film thermally annealed at 600°C with and without prior ion irradiation. The fractal dimension of the percolated pattern in the sample annealed at 600°C was lower than in the sample post-annealed (irradiated and then annealed) at 600°C. All these observations are explained in terms of self-diffusion of Ag atoms on the Si(1 1 1) substrate, inter-diffusion of Ag and Si and phase formations in Ag and Si due to Ar ion irradiation.  相似文献   

7.
The design of a magnetic spectrograph of the QDQ type, to be coupled to a 6 MV Van de Graaff accelerator for the analysis with Rutherford backscattering and recoil spectroscopy of surfaces and thin layers, will be shown. Due to the high energy resolution of. the instrument (ΔE/E ≈ 2 × 10−4) the depth resolution is determined mainly by energy straggling and the combined effect of angular and lateral straggling, and is close to l monolayer near the surface. The mass discrimination of the spectrograph in combination with the energy dispersion of the detector in the focal plane makes possible the background-free detection of light (1H−19F) atoms recoiling from a heavier substrate. The design of the total setup is such that channeling and channeling plus blocking experiments can be carried out, so that the position of light (or heavy) atoms on or in monocrystalline samples can be measured. The detector in the focal plane is position-sensitive in two dimensions so that with one setting of the spectrograph an energy range of 2% and an angular distribution of 5° (with a resolution of 0.1°) can be measured simultaneously. The opening angle in the energy-dispersive direction is 0.1°. The whole spectrograph, including the scattering chamber, can be rotated over 120° with respect to the beam line. For this purpose a bellows construction is made between the beam line and the scattering chamber, permitting the rotation while maintaining a vacuum of ≈ 10−10 Torr in the chamber.  相似文献   

8.
Au+ ion implantation with fluences from 1 × 1014 to 3 × 1016 cm−2 into 12CaO · 7Al2O3 (C12A7) single crystals was carried out at a sample temperature of 600 °C. The implanted sample with the fluence of 1 × 1015 cm−2 exhibited photoluminescence (PL) bands peaking at 3.1 and 2.3 eV at 150 K when excited by He–Cd laser (325 nm). This was the first observation of PL from C12A7. These two PL bands are possibly due to intra-ionic transitions of an Au ion having the electronic configuration of 6s2, judged from their similarities to those reported on Au ions in alkali halides. However, when the concentration of the implanted Au ions exceeded the theoretical maximum value of anions encaged in C12A7 (2.3 × 1021 cm−3), surface plasmon absorption appeared in the optical absorption spectrum, suggesting Au colloids were formed at such high fluences. These observations indicate that negative gold ions are formed in the cages of C12A7 by the Au+ implantation if an appropriate fluence is chosen.  相似文献   

9.
Single-pass flow-through tests were conducted to study the effects of self-radiation damage from alpha decay on dissolution kinetics of three radiation-aged Pu-bearing (1 mass% PuO2) borosilicate glasses over a pH interval of 9–12 at 80–88 °C. The chemical compositions of the glasses were identical except the 239Pu/238Pu isotopic ratio, which was varied to yield accumulated doses of 1.3 × 1016, 2.9 × 1017 and 2.6 × 1018 -decays/g at the time of testing. Release of Al, B, Cs, Na, Si and U to solution increased with increasing pH, whereas Ca, Pu and Sr were invariant over the pH interval. Average dissolution rates, based on B release, were identical within experimental uncertainty for all three glass compositions and increased from 0.17 ± 0.07 at pH(23 °C) 9 to 10.6 ± 2.7 (g/(m2 d1)) at pH(23 °C) 12. Release rates of Pu were 102- to 105-fold slower compared to all other elements and were not affected by isotopic composition, self-radiation damage sustained by the glass, or pH. These data demonstrate that self-radiation damage does not affect glass dissolution rates, despite exposure to internal radiation doses for >20 years.  相似文献   

10.
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   

11.
Compressive creep tests of uranium dicarbide (UC2) have been conducted. The general equation best describing the creep rate over the temperature range 1200–1400°C and over the stress range 2000–15000 psi is represented by the sum of two exponential terms ge =A(σ/E)0.9 exp(−39.6 ± 1.0/RT) + B(σ/E)4.5 exp(−120.6 ± 1.7/RT), where pre-exponential factors are A(σ/E)0.9 = 12.3/h at low stress region (3000 psi) and B(σ/E)4.5 = 3.17 × 1013/h at high stress region (9000 psi), and the activation energy is given in kcal/mol. Each term of this experimental equation indicates that important processes occurring during the steady state creep are grain-boundary diffusion of the Coble model at low stress region and the Weertman dislocation climb model at high stress region. Both mechanisms are related to migration of uranium vacancies.  相似文献   

12.
Fatigue crack growth tests were performed on 2¼Cr–1Mo steel specimens machined from ex-service experimental breeder reactor-II (EBR-II) superheater duplex tubes. The tubes had been metallurgically-bonded with a 100 μm thick Ni layer; the specimens incorporated this bond layer. Fatigue crack growth tests were performed at room temperature in air and at 400 °C in air and humid Ar; cracks were grown at varied levels of constant ΔK. In all conditions the presence of the Ni bond layer was found to result in a net retardation of growth as the crack passed through the layer. The mechanism of retardation was identified as a disruption of crack planarity and uniformity after passing through the porous bond layer. Full crack arrest was only observed in a single test performed at near-threshold ΔK level (12 MPa√m) at 400 °C. In this case the crack tip was blunted by oxidation of the base steel at the steel–nickel interface.  相似文献   

13.
The structure of uranyl ion in 1-butyl-3-methylimidazolium nonafluorobutanesulfonate ionic liquid (BMINfO) has been studied with 1H- and 35Cl-NMR, Raman, and UV-visible spectroscopy. In the 1H-NMR spectrum of the BMINfO solution prepared by dissolving UO2(ClO4)2·5 6H2O, the signal of H2O coordinated to UO22+ was observed at 6.64 ppm at 50°C (free H2O in BMINfO: 3.1 ppm at 50°C), suggesting that the uranyl species exists as the aquo complex, [UO2(H2O)n]2+. The signal of the coordinated H2O disappears with heating at 120°C for 3 h under vacuum. This indicates the dehydration from [UO2(H2O)n]2+. On the other hand, the 35Cl-NMR signal of ClO4 as the counter anion of UO22+ was observed at 1011 ppm (vs. Cl in D2O) regardless of heating. This indicates that no ClO4 ion is in the first coordination sphere of UO22+. Furthermore, the UV-visible absorption spectra showed that the characteristic absorption bands due to UO22+ were sharpened with the dehydration. This means the simplification of the structure around UO22+. These results described above suggest that UO22+ in BMINfO has no ligand in its equatorial plane after the dehydration, i.e. UO22+ exists as a bare cation in this system.  相似文献   

14.
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   

15.
In recent years, single-crystal SiC has become an important electronic material due to its excellent physical and chemical properties. The present paper reports a study of the defect reduction and recrysallisation during annealing of Ge+-implanted 6H-SiC. Implants have been performed at 200 keV with doses of 1 × 1014 and 1 × 1015 cm−2. Furnace annealing has been carried out at temperatures of 500, 950 and 1500°C. Three analytical techniques including Rutherford backscattering spectrometry in conjunction with channelling (RBS/C), positron annihilation spectroscopy (PAS) and cross-sectional transmission electron microscopy (XTEM) have been employed for sample characterisation. It has been shown that damage removal is more complicated than in ion-implanted Si. The recrystallisation of amorphised SiC layers has been found to be unsatisfactory for temperatures up to 1500°C. The use of ion-beam-induced epitaxial crystallisation (IBIEC) has been more successful as lattice regrowth, although still imperfect, has been observed to occur at a temperature as low as 500°C.  相似文献   

16.
The structural and kinetic studies of U(VI) complex with benzamidoxime(Hba) as ligand in CD3COCD3 have been studied by means of 1H and 13C NMR. The Hba molecule was found to coordinate to UO22+ in the form of anionic benzamidoximate (ba), and the number of ba coordinated to UO22+ was determined to be 3 by analyzing the chemical shift of 13C NMR signal for Hba in the presence of UO22+. The exchange rate constants(kex) of ba in [UO2(ba)3] were determined by the NMR line-broadening method. The kinetic parameters were obtained as follows: kex(25°C) = 3.1 × 103s−1, ΔH = 35.8 ± 3.5 KJ mol−1, and ΔS = −65 ± 13.7 J K−1 mol−1. The UV-visible absorption spectra of solutions containing UO22+ and Hba were also measured. The molar extinction coefficient of the complex was found to be extremely large compared with those of UO2(L)52+ (L = unidentate oxygen donor ligands) complexes. This is due to the strong electron withdrawing of UO22+ from Hba and suggests that an interaction between UO22+ and Hba is very strong. Such a high affinity of monomeric amidoxime to UO22+ reasonably explains the high adsorptibility of amidoxime resin to U(VI) species, and is considered to result in the high recovery of U(VI) species from sea water using amidoxime resin.  相似文献   

17.
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   

18.
The vaporization of Li4TiO4 has been studied by a mass spectrometric Knudsen effusion method in the temperature range 1082–1582 K. Identified vapors are Li(g), LiO(g), Li2O(g) and Li3O(g). When the vaporization proceeds, the content of Li2O in the Li4TiO4 sample decreases and the condensed phase of the sample changes to β-Li4TiO4 plus l-Li2TiO3 below 1323 K, to β-Li4TiO4 plus h-Li2TiO3 in the range 1323–1473 K and to h-Li2TiO3 plus liquid above 1473 K. On the basis of the partial pressure data, the enthalpies of formation for β-Li4TiO4 from elements and from constituent oxides have been determined to be ΔHf,298°(β-Li4TiO4,s) = −2247.8 ± 14.3 kJ mol−1 and Δfox,298°(β-Li4TiO4, s) = −107.3 ± 14.3 kJ mol−1, respectively.  相似文献   

19.
The rate of the uranium-water vapour reaction has been measured between 30 and 80°C. The measured reaction rate obeys the rate equation: k = 3.0 × 109r1/2 exp(−15.5 kcal/RT) mg U/cm 2 H = 4.0 × 108r exp(−15.5 kcal/RT) mg weight gain/cm2 h, where r is the fractional relative humidity.

This rate equation agrees remarkably well with the literature equation which was derived from much more limited experimental evidence and so the present equation is preferred.  相似文献   


20.
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   

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