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 62 毫秒
1.
《Journal of Non》2006,352(23-25):2637-2642
We investigated pulsed-X-ray-induced attenuation in two pure-silica-core (PSC) optical fibers with low-chloride (Cl) and different hydroxyl (OH) contents. We measured at room temperature the temporal (100 ms–1000 s) and spectral (0.73–3.1 eV) variation of the induced optical absorption after a 1 MeV X-ray pulse in both low- and high-OH PSC multi-mode fiber samples. A component of the transient loss in the low-OH sample was found to comprise absorption bands at 1.63 and 1.88 eV (760 and 660 nm) arising from self-trapped holes (STHs). Already known to be unstable at room temperature, STHs appear to play a key role in the transient responses of low-OH/low-Cl pure-silica-core fibers in the visible/near-infrared part of the spectrum.  相似文献   

2.
L. Vaccaro  M. Cannas  V. Radzig 《Journal of Non》2009,355(18-21):1020-1023
Two variants of the surface-nonbridging oxygen hole center, (Si–O)3Si–O? and (Si–O)2(H–O)Si–O?, stabilized in porous films of silica nano-particles were investigated by time resolved luminescence excited in the visible and UV spectral range by a tunable laser system. Both defects emit a photoluminescence around 2.0 eV with an excitation spectrum evidencing two maxima at 2.0 and 4.8 eV, this emission decreases by a factor ~2 on increasing the temperature from 8 up to 290 K. However, the different local structure influences the emission lineshape, the quantum yield and the decay lifetime. Such peculiarities are discussed on the basis of the symmetry properties of these defects.  相似文献   

3.
J. Ozdanova  H. Ticha  L. Tichy 《Journal of Non》2009,355(45-47):2318-2322
The glasses representing (Bi2O3)x(WO3)y(TeO2)100?x?y and (PbO)x(WO3)y(TeO2)100?x?y systems were prepared. The dilatometric glass-transition temperatures of examined glass samples were found in the region 383–434 °C, the coefficient of thermal expansion varied from 12 to 16 ppm/°C and the density ranged from 6.302 to 6.808 g/cm3. From the optical transmission measurements of thin glassy bulk samples prepared by a glass blowing, the optical gap values were found in the narrow region 3.21–3.36 eV. For the temperature interval 300–480 K, the values of the temperature coefficient of the optical band gap varied from 3.7 × 10?4 to 5.24 × 10?4 eV/K. It is suggested that Raman feature observed at around 350 cm?1 can be assigned to an overlap of Raman bands attributed to WO6 corner shared octahedra and to the following three atomic linkages: Bi–O–Te, Pb–O–Te and W–O–Te.  相似文献   

4.
Doris Ehrt 《Journal of Non》2008,354(2-9):546-552
Glasses with 55–60 mol% SnO and 40–45 mol% P2O5 have shown extremely large differences in the chemical and thermal properties depending on the temperature at which they were melted. Glasses prepared at low melting temperature, 450–550 °C, had low Tg, 150–200 °C, and low chemical stability. Glasses prepared at high melting temperature, 800–1200 °C, had much higher Tg, 250–300 °C, and much higher chemical stability. No significant differences were found by 119Sn Mössbauer and 31P Nuclear Magnetic Resonance spectroscopy. Large differences in the OH-content could be detected as the reason by infrared absorption spectroscopy, thermal analyses, and 1H Nuclear Magnetic Resonance spectroscopy. In samples with low Tg, a broad OH – vibration band around 3000 nm with an absorption intensity >20 cm?1, bands at 2140 nm with intensity ~5 cm?1, at 2038 nm with intensity ~2.7 cm?1, and at 1564 nm with intensity ~0.4 cm?1 were measured. These samples have shown a mass loss of 3–4 wt% by thermal gravimetric analyses under argon in the temperature range 400–1000 °C. No mass loss and only one broad OH-band with a maximum at 3150 nm and low absorption intensity <4 cm?1 could be detected in samples melted at high temperature, 1000–1200 °C, which have much higher Tg, ~300 °C, and much higher chemical stability.  相似文献   

5.
Corrosion of nuclear waste glass in unsaturated conditions is expected to occur upon the closure of the repository galleries during disposal cell saturation in the proposed French disposal site. The objectives of the present work were to determine the alteration kinetics of the SON68 reference in such conditions. Vapor hydration tests were conducted using thin, polished SON68 glass coupons contained in stainless steel autoclaves. Temperatures ranged between 90 °C and 200 °C and the relative humidity (RH) was maintained at 91 ± 1%. Additional experiments at 175 °C and 80, 85, 90 and 95% RH were also conducted to assess the role of RH on the glass corrosion rate. The nature and extent of corrosion have been determined by characterizing the reacted glass surface with scanning electron microscopy (SEM), transmission electron microscopy (TEM), and energy dispersive X-ray spectroscopy (EDS). Elemental profiling of the glass hydrated at 90 °C was studied by TOF-SIMS. The chemical composition of the external layer depends on experimental conditions. The hydration rate at 90 °C (TOF-SIMS analysis) is 10 × higher than the generally accepted final rate of SON68 in water at 90 °C (~ 10? 4 g m? 2 d? 1). This may indicate that the glass hydration process cannot be simulated by experiments in aqueous solution with a high S/V ratio. Subsequent leaching (corrosion in an aqueous solution) of samples weathered in water vapor showed dissolution rate values higher than those of pristine glass. This result indicates that mobile elements are trapped within the alteration products during the hydration step and it gives insight into mobility variations of the considered elements.  相似文献   

6.
《Journal of Non》2006,352(26-27):2917-2920
Our previous studies have reported the excitation energy dependence of the 2.7 and 4.3 eV photoluminescence (PL) bands in oxygen deficient silica glass at low temperature (∼20 K). An oxygen vacancy (O3SiSiO3) was thought to be the origin of the two PL bands. In order to verify the origin of the 2.7 and 4.3 eV PL bands in silica glass, we measured the PL band of various thermally heat treated silica glasses. In the sample after heat treatment, we did not observe the 4.3 eV PL band, though we did observe the 2.7 eV PL band. These results suggest that these two PL bands do not have a common origin.  相似文献   

7.
Fast ion conducting (FIC) phosphate glasses and glass ceramic composites have gained considerable importance due to their potential applications in the fabrication of solid-state batteries and other electrochemical devices. We, therefore, present an overview on various types of FIC glasses and glass ceramic composites. Silver phosphate glasses doped with different weight percent of lithium chloride (1, 5, 10 and 15 wt.%) were synthesized by melt quenching technique. The Ag2O–P2O5–(15 wt.%) LiCl glass exhibited the maximum electrical conductivity (σ = 8.91 × 10? 5 S cm? 1 at room temperature and 4.16 × 10? 3 S cm? 1 at 200 °C). Using this glass as an amorphous host material, glass–ceramic composites of Ag2O–P2O5–(15 wt.%) LiCl:xAl2O3 (x = 5–50 wt.%) were prepared. The ionic transference number, electrical conductivity, ionic mobility and carrier ion concentration of the synthesized samples were measured. Ag2O–P2O5–(15 wt.%) LiCl:(25 wt.%) Al2O3 composite system exhibited the maximum σ value (σ = 3.32 × 10? 4 S cm? 1 at room temperature and 2.88 × 10? 2 S cm? 1 at 200 °C ). Solid‐state batteries using undoped Ag2O–P2O5 glass, Ag2O–P2O5–(15 wt.%) LiCl glass and glass ceramic composite containing 25 wt.% Al2O3 as electrolytes were fabricated. The open circuit voltage (OCV) values and discharge time of these cells were measured and compared. It is found that the glass ceramic composites show enhanced ionic conduction, better OCV value and discharge characteristics.  相似文献   

8.
《Journal of Non》2005,351(40-42):3334-3340
We have measured and analyzed the optical constants and polarized optical properties of amorphous aluminum nitride (a-AlN) thin films deposited by RF reactive magnetron sputtering onto silicon(1 1 1) and glass substrates. The optical constants were obtained by analysis of the measured ellipsometric spectra in the wavelength range 300–1400 nm, using the Cauchy–Urbach model. Refractive indices and extinction coefficients of the films were determined to be in the range 1.80–2.11 and 8.6 × 10−3–1.5 × 10−5, respectively. Analysis of the absorption coefficient, in the wavelength range 200–1400 nm, shows the bandgap of a-AlN thin films to be 5.84 ± 0.05 eV. From the angle dependence of the p-polarized reflectivity we deduce a Brewster angle of 61° and a principal angle of 64°. Measurement of the polarized optical properties reveals a high transmissivity and very low absorptivity for a-AlN films in the visible and near infrared regions. X-ray diffraction analysis confirmed the amorphous nature of the films under study.  相似文献   

9.
We present transmission electron microscopy and Raman scattering measurements showing that niobium inhibits the processes of nucleation and growth of anatase crystallites in the initial amorphous titania nanotubes and thus shifts the temperature of the complete amorphous-to-anatase phase transition to higher values up to 550 °C. Niobium dopants stabilize the anatase phase in titania nanotubes up to 650 °C. The size of anatase crystallites can reach 30–50 nm. Excess niobium atoms which are pulled off from the volume of anatase crystallite form polymeric or monomeric Ti–O–NbO groups at the interface area. Slight shift and broadening of Eg (144 cm?1), A1g (515 cm?1) and Eg (630 cm?1) modes in Raman spectra can be explained by niobium insertion into the anatase structure.  相似文献   

10.
Amorphous silicon quantum dots (Si-QDs) self-aggregated in silicon-rich silicon carbide are synthesized by growing with plasma-enhanced chemical vapor deposition on (100)-oriented Si substrate. Under the environment of Argon (Ar)-diluted Silane (SiH4) and pure methane (CH4), the substrate temperature and RF power are set as 350 °C and 120 W, respectively, to provide the Si-rich SiC with changing fluence ratio (R = [CH4 ]/[SiH4] + [CH4]). By tuning the fluence ratio from 50% to 70%, the composition ratio x of Si-rich Si1 ? xCx film is varied from 0.27 to 0.34 as characterized by X-ray photoelectron spectroscopy (XPS), which reveals the component of Si2p decreasing from 66.3 to 59.5%, and the component of C1s increasing from 23.9% to 31% to confirm the formation of Si-rich SiC matrix. Annealing of the SiC sample from 650 °C to 1050 °C at 200 °C increment for 30 min induces the very tiny shift on the wavenumber of the crystalline Si (c-Si) related peak due to the precipitation of Si-QDs within the SiC matrix, and the Raman scattering spectra indicate a broadened Raman peak ranging from 410 to 520 cm? 1 related to the amorphous Si accompanied with the significant enhancement for SiC bond related peak at 980 cm? 1. From the high resolution transmission electron microscopy images, the critical temperature for Si-QD precipitation is found to be 850 °C. The self-assembly of the crystallized Si-QDs with the size of 3 ± 0.5 nm and the volume density of (3 ± 1) × 1018 (#/cm3) in Si-rich SiC film with R = 70% are observed after annealing at higher temperature.  相似文献   

11.
Ca-chloroapatite (CaApCl), glass-bonded CaApCl compositions loaded with 16–32 wt.% simulated pyrochemical chloride waste were prepared by mixing and heating (773–1023 K) apatite and borosilicate glass (BSG) forming reagents in appropriate ratios. The compositions were characterized by XRD, TGA/DTA, SEM, and EDAX. Among the products, 16–27 wt.% chloride waste loaded composition yielded phase pure Ca-chloroapatite and were resistant to leaching of Cl? and other ions. In case of 28–32 wt.% waste loaded compositions, even though formation of phase pure Ca-Chloroapatites was observed by XRD, the leaching of Cl? and other ions was found to be significant. Bulk thermal expansion behavior of the samples was studied by dilatometry. The 16 wt.% chloride waste loaded matrix showed nearly the same thermal expansion compared to pure Ca-Chloroapatites. The % linear thermal expansion of the matrices decrease on increasing the chloride waste loading; however, Ca-chloroapatite mixed with 20 wt.% BSG mixed matrix showed slightly higher thermal expansion. The coefficient of thermal expansion of borosilicate glass is the lowest among all the matrices measured. The coefficient of thermal expansion (CTE) is found to be 12.76 ± 0.64 × 10? 6 K? 1 for CaApl and 12.18 ± 0.63 × 10? 6 K? 1 for 16 wt.% waste loaded BSG-encapsulated CaApl in the temperature range of 298–780 K. The glass transition temperature of the waste loaded matrices is lower than that of the bare BSG and 20 wt.% BSG encapsulated Ca-chloroapatite.  相似文献   

12.
Estimates of Kerr electrooptical sensitivity of several tellurite glasses are presented. The highest value of Kerr coefficient B  190 × 10?16 m V?2 is registered for 0.6TeO2–0.3TlO0.5–0.1ZnO glass. This evidences the prospects of thallium–tellurite glass system for electrooptical applications. A gradual decrease of B from 41 × 10?16 to 26 × 10?16 m V?2 in (1 ? x) TeO2  xNbO2.5 system is revealed for x increasing from 0.1 to 0.15. No crystalline phase was found in that system, thus allowing attributing its Kerr sensitivity to the intrinsic properties of the glass matrix. The Kerr coefficient variation from 66 to 81 × 10?16 m V?2 was observed for 0.85TeO2–0.15WO3 glasses co-doped with small amounts of silver and cerium. The analysis of optical absorption spectra of several silver-containing tellurium–tungsten oxide glasses makes it possible to think that introducing cerium provokes formation of new mid-range orderings.  相似文献   

13.
We have studied the impact of temperature and pressure on the structural and electronic properties of Ge:P layers grown with GeH4+PH3 on thick Ge buffers, themselves on Si(0 0 1). The maximum phosphorous atomic concentration [P] exponentially decreased as the growth temperature increased, irrespective of pressure (20 Torr, 100 Torr or 250 Torr). The highest values were however achieved at 100 Torr (3.6×1020 cm?3 at 400 °C, 2.5×1019 cm?3 at 600 °C and 1019 cm?3 at 750 °C). P atomic depth profiles, “box-like” at 400 °C, became trapezoidal at 600 °C and 750 °C, most likely because of surface segregation. The increase at 100 Torr of [P] with the PH3 mass-flow, almost linear at 400 °C, saturated quite rapidly at much lower values at 600 °C and 750 °C. Adding PH3 had however almost no impact on the Ge growth rate (be it at 400 °C or 750 °C). A growth temperature of 400 °C yielded Ge:P layers tensily-strained on the Ge buffers underneath, with a very high concentration of substitutional P atoms (5.4×1020 cm?3). Such layers were however rough and of rather low crystalline quality in X-ray Diffraction. Ge:P layers grown at 600 °C and 750 °C had the same lattice parameter and smooth surface morphology as the Ge:B buffers underneath, most likely because of lower P atomic concentrations (2.5×1019 cm?3 and 1019 cm?3, respectively). Four point probe measurements showed that almost all P atoms were electrically active at 600 °C and 750 °C (1/4th at 400 °C). Finally, room temperature photoluminescence measurements confirmed that high temperature Ge:P layers were of high optical quality, with a direct bandgap peak either slightly less intense (750 °C) or more intense (600 °C) than similar thickness intrinsic Ge layers. In contrast, highly phosphorous-doped Ge layers grown at 400 °C were of poor optical quality, in line with structural and electrical results.  相似文献   

14.
In order to crystallize a large quantity of the lithium?mica in glass?ceramics, 5.1 mass% MgF2 was added to the starting materials of the parent glasses having chemical compositions of Li(1+x)Mg3AlSi3(1+x)O10+6.5xF2 (x = 0.5 and 1.0). Transparent glass?ceramics, in which a large quantity of lithium?mica with particle size of <50 nm was separated, could be prepared from the MgF2-added parent glass with x = 0.5. While the parent glass, which had a binodal phase separation structure, did not exhibit electrical conductivity, the transparent glass–ceramic was given conductivity by the formation of an interlocking structure of mica. As the separated mica formed a tighter interlocking structure, the conductivity increased and reached a value of 2.0 × 10?3 S/cm at 600 °C. The MgF2-added parent glass with x = 1.0 was not transparent because of coarse spinodal phase separation. The conductivity was 4.3 × 10?4 S/cm at 600 °C but was significantly decreased by the separation of mica.  相似文献   

15.
Efficient infrared emissions at 1.20 μm [5I6  5I8 transition] and 1.38 μm [(5 F4, 5 S2)  5I5 transition] from Ho3+-doped lithium–barium–bismuth–lead (LBBP) glass were observed. The stimulated emission cross-sections were calculated to be 0.29 × 10?20 and 0.25 × 10?20 cm2 for 1.20 and 1.38 μm emissions, respectively. Judd-Ofelt characteristic parameters Ω2, Ω4 and Ω6 for Ho3+ in LBBP glass were calculated to be 6.72 × 10?20, 2.35 × 10?20 and 0.61 × 10?20 cm2, respectively, which indicates a strong asymmetry and a covalent environment between the Ho3+ ions and the ligands in this glass. The optical amplifications operating at these relatively unexplored wavelength regions were evaluated and discussed.  相似文献   

16.
Glasses in the system MgO/Al2O3/TiO2/ZrO2/SiO2 with and without the addition of As2O3 and Sb2O3 were thermally treated. Up to a temperature of 950 °C, this resulted in the formation of ZrTiO4, sapphirine and high quartz solid solution. Annealing at higher temperatures led to the formation of low quartz solid solutions, ZrTiO4 and sapphirine. This resulted in a continuous increase of density, hardness, fracture toughness and thermal expansivity. In the glass doped with As2O5 and Sb2O5 annealing temperatures >1000 °C resulted in the formation of cristobalite instead of quartz. Then the density, hardness and strength decreased again, while the fracture toughness (up to 2.8 MPa m1/2) and the thermal expansion coefficient increased strongly. In the dilatometric curves, a steep increase in expansion is observed in the temperature range from 100 to 200 °C, which is attributed to the transformation of low cristobalite to high cristobalite. The mean linear thermal expansion coefficient (25–200 °C) is 20 × 10?6 K?1 and the largest up to now reported in the literature for alkali-free silicate glass–ceramics.  相似文献   

17.
The grain boundary groove shapes for equilibrated solid neopentylglycol (NPG) solution (NPG–3 mol% D-camphor) in equilibrium with the NPG–DC eutectic liquid (NPG–36.1 mol% D-camphor) have been directly observed using a horizontal linear temperature gradient apparatus. From the observed grain boundary groove shapes, the Gibbs–Thomson coefficient (Г), solid–liquid interfacial energy (σSL) of NPG solid solution have been determined to be (7.5±0.7)×10?8 K m and (8.1±1.2)×10?3 J m?2, respectively. The Gibbs–Thomson coefficient versus TmΩ1/3, where Ω is the volume per atom was also plotted by linear regression for some organic transparent materials and the average value of coefficient (τ) for nonmetallic materials was obtained to be 0.32 from graph of the Gibbs–Thomson coefficient versus TmΩ1/3. The grain boundary energy of solid NPG solution phase has been determined to be (14.6±2.3)×10?3 J m?2 from the observed grain boundary groove shapes. The ratio of thermal conductivity of equilibrated eutectic liquid to thermal conductivity of solid NPG solution was also measured to be 0.80.  相似文献   

18.
Routes to atomic layer-deposited TiO2 films with decreased leakage have been studied by using electrical characterization techniques. The combination of post-deposition annealing parameters, time and temperature, which provides measurable aluminum–titanium oxide–silicon structures – i.e., having capacitance–voltage curves which show accumulation behavior – are 625 °C, 10 min for p-type substrates, and 550 °C, 10 min for n-type substrates. The best annealing conditions for p-type substrates are 625 °C with the length extended to 30 min, which produces an interfacial state density of about 5–6 × 1011 cm?2 eV?1, and disordered-induced gap state density below our experimental limits. We have also proved that a post-deposition annealing must be applied to TiO2/HfO2 and HfO2/TiO2/HfO2 stacked structures to obtain adequate measurability conditions.  相似文献   

19.
A glass of composition (20 ? x)Li2O–xLiCl–65B2O3–10SiO2–5Al2O3 where 0 ? x ? 12.5 wt% is prepared using the normal melt-quenching technique. The optical constants and electrical conductivity and their correlation are investigated, furnished and discussed with the substitution of Li2O for LiCl. The mechanism of the optical absorption and the calculated Urbach energy follow the rule of phonon-assisted transitions. The ionic conduction mechanism is determined by activation energy process. Substitution up to 10 wt% LiCl provides high ionic conductivity (1.9 × 10?2 Ω?1 m?1) due to the high average electronegativity of LiCl which increases the polarizability of lithium ions. The small cation–anion distance approach confirmed the enhancement in ionic conductivity of LiCl containing glass compared to that of Li2O. Due to the large size of Cl? ions, there is an expansion of the lattice which in turn broadens the available path windows. For 12.5 wt% LiCl, anomalous density behavior is observed and a reduction in conductivity is occurred, σ = 5.4 × 10?3 Ω?1 m?1. Owing to the model of bond fluctuation, the reduction is attributed to the increase in the alkali halide concentration which creates bottlenecks that hinder the motion of Li+ ions. The ionic conductivity character is strongly supported by the behavior of the glass ionicity factor, density, molar volume, refractive index, average boron–boron separation, molar refraction, metallization criterion and non-bridging oxygen concentration of the studied glass.  相似文献   

20.
《Journal of Non》2007,353(52-54):4743-4752
Cation diffusion was experimentally investigated in soda-lime-silicate glass melts (composition in mol%: 74SiO2–16Na2O–10CaO) at temperatures from 1000 to 1200 °C using the diffusion couple technique. One half of each diffusion couple was doped with 11 trace elements (500 ppm by weight of Rb, Cs, Sr, Zn, Cd, Nd, Eu, In, Sn, Ge and 1000 ppm by weight of Fe). Experiments were performed in an internally heated gas pressure vessel at a confining pressure of 100 MPa to avoid convective fluxes in the diffusion samples. The distribution of major elements was analyzed by electron microprobe. IR spectroscopy was used to quantify concentrations of dissolved water in the run products. Trace element diffusion profiles were measured simultaneously employing synchrotron X-ray fluorescence microanalysis. In all analyzed glasses the highest diffusion coefficients were observed for Rb whereas Nd was always the slowest element, e.g. at 1000 °C the diffusivity decreases from (1.51 ± 0.35) × 10−11 m2/s for Rb to (1.29 ± 0.34) × 10−13 m2/s for Nd. The diffusivity of Nd is close to the chemical diffusivity of network former calculated from viscosity data using the Eyring relationship. Surprisingly, the rare earth elements Nd (3+) and Eu (mixed 2+, 3+) diffuse more slowly than the tetravalent Ge. Activation energies for diffusion increase from (132.1 ± 1.5) kJ/mol for Rb to (205 ± 16) kJ/mol for Eu. Based on the diffusion data for Eu, Sr and Nd we estimated that Eu2+/Eutotal ratios in soda-lime-silicate glass melts are below 0.04 both at reducing and oxidizing conditions.  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