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 behavior of copper ions in the Cu2O·Al2O3·4SiO2 (in moles) glass on heating in air at temperatures up to 500°C was studied. When the glass, in which about 90% of Cu was present as Cu+ ions, was heated in air above 300°C, a CuO layer was formed on the surface. The amount of CuO was increased with heating temperature and time, corresponding to the decrease in weight of the glass. Furthermore, the fraction of Cu2+ ions in the glass increased. These observations suggest that oxygens do not diffuse into the glass, but Cu+ ions migrate to the surface from the interior to balance the surplus positive charge produced by the oxidation of Cu+ to Cu2+ ions inside the glass. The following reaction scheme for the formation of the CuO layer was proposed; 2Cu+(interior) + 21O2(surface) → Cu2+(interior) + CuO(surface).  相似文献   

2.
Diffusion coefficients of iron were measured in glass melts with the basic compositions 5Na2O · xMgO · (15−x)CaO · yAl2O3 · (80−y)SiO2 with x=5, 10 and y=0, 5, 7.5, 10 and 15. The melts were doped with 0.25 mol% Fe2O3 and studied in the temperature range from 1000 to 1600 °C using square-wave voltammetry. The voltammograms exhibited distinct peaks attributed to the reduction of Fe3+ to Fe2+, from which peak currents mixed diffusion coefficients of iron were calculated. Diffusion coefficients in all melt compositions which did not show crystallization could be fitted to Arrhenius equation. The diffusivities measured in different melt compositions were related to the same viscosity, i.e. not the same temperature. Increasing the alumina concentration from 5 to 10 mol% resulted in an increase of the viscosity corrected diffusivities. At further increasing alumina concentrations, the diffusivities get smaller again. This can be explained by the stabilizing effect of Na+ and Ca2+ on FeO4 and AlO4-tetrahedra, which strengthens the incorporation of Fe3+ into the glass structure.  相似文献   

3.
Thermally stimulated luminescence (TSL) and infrared (IR) spectroscopy were measured in plasma grown Si1−xGexO2 (x=0, 0.08, 0.15, 0.25, 0.5) with different thicknesses (12–40 nm). A comparison with the TSL properties of thermally grown SiO2 and GeO2 was also performed. A main IR absorption structure was detected, due to the superposition of the peaks related to the asymmetric O stretching modes of (i) Si–O–Si (at ≈1060 cm−1) and (ii) Si–O–Ge (at 1001 cm−1). Another peak at ≈860 cm−1 was observed only for Ge concentrations, x>0.15, corresponding to the asymmetric O stretching mode in Ge–O–Ge bonds. A TSL peak was observed at 70°C, and a smaller structure at around 200°C. The 70°C peak was more intense in all Ge rich layers than in plasma grown SiO2. Based on the thickness dependence of the signal intensity we propose that at Ge concentrations 0.25x0.5 TSL active defects are localised at interfacial regions (oxide/semiconductor, Ge poor/Ge rich internal interface, oxide external surface/atmosphere). Based on similarities between TSL glow curves in plasma grown Si1−xGexO2, thermally grown GeO2 and SiO2 we propose that oxygen vacancy related defects are trapping states in Si1−xGexO2 and GeO2.  相似文献   

4.
Ag+/Na+ ion-exchanged R2O–Al2O3–SiO2 glasses with uniform concentration profile of Ag+ and Na+ were prepared by heat treatment in molten silver salt followed by holding at the same temperature in an ambient atmosphere. Their glass transition temperature (Tg) and thermal expansion coefficient (TEC) were measured and structures were investigated using 29Si-MAS NMR, 27Al-MAS NMR, IR and Raman spectroscopies. Both Tg and TEC decreased with increase of the exchange ratio, but Tg was still above the ion-exchange temperature of 400°C even for the fully exchanged sample. The 29Si- and 27Al-MAS NMR spectra were mostly unchanged and no sign of the structural alteration of the glass network was observed. On the other hand, the vibrational spectra showed remarkable peak shifts depending on the exchange ratio. From these structural results, it was found that when the exchange ratio was low, the introduced Ag+ ions were stabilized at the non-bridging oxygen (NBO) site, and then Na+ ions in AlØ4 site were exchanged by Ag+ ions after full replacement of NBO sites, where Ø represents the bridging oxygen.  相似文献   

5.
A series of titania-silica glasses with 0–9% TiO2 were fabricated using a sol/gel process. The sol was prepared by dispersing colloidal silica fume in an aqueous solution of titania which was synthesized through the acid-catalyzed hydrolysis of titanium isopropoxide. The sols gelled in 2–4 days, and then were dried for 6–8 days. The dry gels were sintered at 1450–1500°C to produce clear, dense, microstructure-free glasses. The gels underwent a total shrinkage of 50% to yield glass rods about 50 mm long and 5 mm in diameter, or glass discs about 4 cm in diameter and 5 mm thick. The drying step was most critical in the production of crack-free specimens.

In the gel, the transmission electron microscope (TEM) revealed the presence of 1–5 nm rutile microcrystallites uniformly distributed within a network of colloidal silica particles. After sintering to 1450–1500°C, though, a dense, transparent, microstructure-free glass was created. Fourier transform infrared spectroscopy (FTIR) verified the formation of an amorphous solid-solution of titania and silica after sintering.

The thermal expansion of the glasses was measured using a differential dilatometer. The average linear coefficients of thermal expansion (CTE @ 25–675°C) varied between +5 × 10−7 and −0.2 × 10−7°C−1 in the range 0 to 9% TiO2. The glass with 7.2% TiO2 exhibited a zero thermal expansion coefficient at 150–210°C. The hysteresis in CTE on heating and cooling was of the order of 0.01–0.02 ppm.  相似文献   


6.
K. Hirao  T. Komatsu  N. Soga 《Journal of Non》1980,40(1-3):315-323
Mössbauer absorption measurements have been made at room temperature on 57Fe in iron sodium silicate glasses containing 3–15 mol% Fe2O3 and various iron alkali silicate crystals in order to study the state of iron in these glasses. The spectra of all the glasses gave one doublet with a quadrupole splitting varying from 0.73–0.78 mm s−1, while those of Na2O · Fe2O3 · 4 SiO2 and 5 Na2O · Fe2O3 · 8 SiO2 crystals showed much smaller quadrupole splitting, 0.28 mm s−1 and 0.10 mm s−1, respectively, and an asymmetrical doublet of much narrower linewidth. When sodium was replaced by other alkali metals of larger size, such as K and Cs, in MFeSi2O6 and MFeSi3O8 crystals, the quadrupole splitting became wider and approached to 0.73 mm s−1. Such a variation was not observed for glasses. These results suggest that a larger number of non-identical sites exist in iron sodium silicate glasses than in the corresponding crystals.  相似文献   

7.
The gel formation of the (100-x)TiO2·xSiO2 (x = 0–10 mol%) system has been studied. The progressive elimination of residues was followed by DTA and TGA curves. DTA curves showed that the formation of anatase during heat treatment could be sensibly slowed down with the increase of SiO2. The relationship between the gel composition and crystallization temperature of anatase has been systematically investigated. The X-ray diffraction spectra demonstrated that the crystallization temperature of anatase is 400°C for TiO2 gel and 430°C for 90TiO2 - 10SiO2 gel. The infrared absorption spectra were used to follow the structural transformation of gels heat-treated at different temperatures. With the help of EPR it is evident that titanium ions exist only in tetravalence.  相似文献   

8.
Glasses of compositions 5ZrO2·5SiO2(ZS), 5ZrO2·Al2O3·4SiO2(ZAS) and 5 5ZrO2·0.5Al2O3·0.5Na2O·4SiO2(ZANS) were prepared by the sol-gel process from metal alkoxides and sintered to make glass-ceramics. Tetragonal ZrO2 was precipitated by heat treatment at 900 to 1300°C. The activation energy for tetragonal ZrO2 crystal growth was extremely high in Al2O3 containing glasses. ZAS and ZS were sintered to the near theoretical densities above 1200°C, at which the predominant phase was tetragonal ZrO2. On the other hand, for ZANS, high densification was not attained owing to the large pores enclosed by the glass phase. Strength and fracture toughness increased with the densification and the crystal growth of tetragonal ZrO2, reaching 450 MPa and 9 MN/m1.5, respectively.  相似文献   

9.
Raman spectra have been measured for ZnCl2---ZnX2 and ZnCl2---KX (X = Br, I) glasses to investigate the structure of the glasses with varying composition. The assignment of each band was made, and the change of the spectra with composition was explained in terms of the bridging and non-bridging states of halide ions and the change of the tetrahedral units, ZnXnCl4−n2− (n = 0–4), formed in the glasses. As the content of ZnX2 in ZnCl2---ZnX2 glasses increases (20 → 80 mol%), the peak frequency of the Zn---Cl stretching mode increases (238 → 248 cm−1 in X = I glasses, 238 → 259 cm−1 in X = Br glasses) while the Zn---I and Zn---Br stretching frequencies decrease (173 → 120 cm−1 for Zn---I, 196 → 157 cm−1 for Zn---Br). The decrease of the Zn---I and Zn---Br band frequencies was attributed to the increase of the number n of the ZnXnCl4−n2− tetrahedra. The increase of the Zn---Cl frequency suggests the existence of the bonding state of Cl ions which is intermediate between the bridging and the non-bridging states. In ZnCl2---KX glasses, the Zn---Clnon-bridging band at about 300 cm−1 was observed in addition to the bands observed in ZnCl2---ZnX2 glasses. The addition of KX produces non-bridging anions while the tetrahedral units, ZnXnCl4−n2− are also formed.  相似文献   

10.
The La L1 and L3 XANES and L3 EXAFS have been investigated for the series of glasses 10K2O---50SiO2---x La2O3 (x = 1, 5, 10) and (10 − x)K2O---40SiO2−(x/3)La2O3 (x = 7.5, 5, 2.5) and model compounds La2O3, LaAlO3, LaPO4, La2NiO4, La2CuO4 and La(OH)3. An edge resonance at 25 eV above the L1 edge in the glass spectra is concentration-dependent, decreasing in intensity with increasing lanthanum concentration. The 2s → nd forbidden transition increases with La2O3 concentration, indicating a reduction in the ‘average’ site symmetry of the first coordination shell of La. Mapping X(k) space, which is a new and promising technique, was employed to extract bond distance, coordination number and thermal parameters from the EXAFS. By this method, one calculates the complete X(k) space a function of all physically reasonable values of the adjusted parameters in all possible combinations. The advantage in this method is the assurance of a global minimum. Bond lengths were comparable to those obtained by Fourier transforming the phase corrected EXAFS. The values are 2.42 Å (± 0.03 Å) for La---O. The coordination numbers (N ≤ 7 ± 1.5) were derived by mapping and comparison to the published structures for other La compounds. X(k) mapping is compared with least-squares fitting the data, and the correlation between the Debye-Waller factor and coordination number is also discussed.  相似文献   

11.
The electrical conductivities of (1−x) Li2O · x BaO · 2 SiO2, (1−x) Na2O · x MgO ·2 SiO2, (1−x) Na2O · x CaO · SiO2 and (1−x) Na2O · x BaO · 2SiO2 glasses were measured at temperature ranging from room temperature to 450°C. The transport numbers for Na+ ion in (1−x) Na2O · x BaO · 2 SiO2 glasses were measured. It was found that the alkali ion carried a significant part of the current in these glasses except one that had no alkali ions, and the conductivity decreased markedly as the alkali oxide was substituted by an alkaline earth oxide. The results of conductivity measurements combined with the data hitherto reported on mixed alkali glasses led to the proposal that the so-called mixed alkali effect could be explained on the basis of the independent path model in which it is assumed that cations can move only through vacant sites left by those of the same type.  相似文献   

12.
Various oxide films on SiO2 glass substrates were irradiated by a laser beam. A continuous CO2 laser source (wavelength 10.6 μm) was used for this purpose; the composition change at the surface layer was determined by Rutherford backscattering spectrometry (RBS). All the alkaline-earth oxides as well as those of lanthanum and yttrium, entered the glass after treatment. ZrO2 and CeO2, however, did not enter the SiO2 glass due to laser irradiation. It is interesting, however, that a film of ZrO2 + Al2O mixture easily entered into the SiO2 glass by laser processing. The conditions and mechanism of laser-enhanced interaction of ZrO2 or other oxide films with SiO2 glass surfaces are discussed especially in view of their structural behaviour in glass.  相似文献   

13.
High optical quality Er3+:YVO4 laser crystals have been grown by using the floating zone method (FZ). The spectroscopic properties and 3 μm lasing of Er3+:YVO4 were investigated. It is found that the Er3+ concentration has a negative effect on the emission of the transition 4I13/24I15/2(1.55 μm ), and a positive effect on that of the 4I11/24I13/2 transition (2.68 μm). With direct upper-state pumping and a plane-concave cavity a self-terminating laser was achieved at the wavelength of 2.724 μm in the 30 at% Er3+ doped sample.  相似文献   

14.
The effect of annealing on the structure of Ge20Te80 glass has been examined in atomic-scale images obtained using a scanning tunnelling microscope (STM). The STM image has been able in part to answer the question as to how the atomic structure is changed by low-temperature (<Tg) annealing, whereas the effect of annealing on the structure was never observed in previous neutron diffraction study. Scanning tunneling microscope images as large as ≈ 100 nm2 have provided atomic-resolution ridges ≈ 3 nm in length. The nearest neighbour distance between peaks in each alignment is equal to ≈ 0.5 nm; the alignments in parallel with each other are a distance of ≈ 0.7 nm apart. These ridges can generate a surface associated with pseudo flatness, the size of which is less than ≈ 10 nm2. Thus, the effect of annealing on the structure of as-quenched Ge20Te80 glass has given rise to intermediate-range order in a region less than ≈ 10 nm2. The atomic configuration of annealed Ge20Te80 glass is inhomogeneous in the range < 10 nm2.  相似文献   

15.
J. Gtz 《Journal of Non》1976,20(3):413-425
The type and the amount of silicate groupings existing in glassy and crystalline 2PbO·SiO2 have been determined by direct chemical methods: paper chromatography, trimethylsilylation combined with gas-liquid partition chromatography and by the molybdate method. The results obtained by these three different methods are in good agreement and demonstrate, that glassy 2PbO·SiO2 and each of the three main crystalline polymorphs are characterized by its own specific silicate anion distribution: the distribution in vitreous 2PbO·SiO2 is of a polyanionic nature; in T---Pb2SiO4 dimetic groups [Si2O7]6− prevail; M1---Pb2SiO4 contains predominantly [Si4O12]8− rings and H---Pb2SiO4 is a typical polysilicate with chain anions [SiO32−]n. The results fit a structural model according to which glass is a random array of discrete polyatomic groupings; the gradual transition from the glassy state to the stablest crystalline structure is connected with degradation and polymerization of silicate anions.  相似文献   

16.
E Prasad  M Sayer  H.M Vyas 《Journal of Non》1980,40(1-3):119-134
Glasses of composition 65 mol% LiNbO3:: 35 mol% SiO2 have been shown to be Li+ ion conductors with a conductivity at 200°C > 1 × 10−5 (η cm)−1 and an activation energy of 0.54 eV. The addition of approximately 0.1 mol% Fe2O3 leads to an enhancement of conductivity to ≈10−3 (η cm)−1 at 200°C and an activation energy of 0.67 eV. The effect of Fe is shown to be in the control of microstructure in the glass, with Fe2O3 concentrations < 1 mol% acting as a grain growth inhibitors and larger concentrations acting as a nucleating agents. A model for this process based on the expected stoichiometry of the melt and the effect of Fe2+ and Fe3+ in charge compensation is in excellent agreement with experimental data from electron spin resonance.  相似文献   

17.
《Journal of Non》2001,290(2-3):224-230
Infrared (IR) reflectivity and Raman scattering spectra of LaBSiO5 glass and glass–crystal composites were studied in the temperature range 25–260 °C. Using an analogy with the LaBGeO5 crystal it was possible to assign the main spectral features of the LaBSiO5 glass and glass–crystal composites. The BO4 chain arrangement and the bending vibrations of SiO4 are influenced by the loss of the long-range order in the glass whereas the stretching vibrations of the SiO4 groups are practically unaffected. The structural disorder in LaBSiO5 crystallites is caused by rotation of BO4 tetrahedra.  相似文献   

18.
Fine-sized ZnO–B2O3–CaO–Na2O–P2O5 glass powders with spherical shape were directly prepared by high temperature spray pyrolysis. The ZnO–B2O3–CaO–Na2O–P2O5 powders prepared by spray pyrolysis at temperatures above 1200 °C had broad peaks at around 30° in the XRD patterns. The glass transition temperatures (Tg) of the glass powders obtained by spray pyrolysis at preparation temperatures between 900 °C and 1400 °C were near 480 °C regardless of the preparation temperatures. The dielectric layers formed from the glass powders prepared by spray pyrolysis at preparation temperatures above 1300 °C had clean surface and dense inner structure at the firing temperature of 580 °C. The transmittance of the dielectric layer formed from the glass powders obtained by spray pyrolysis at preparation temperature of 1400 °C was 90% at the firing temperature of 580 °C, in which the thickness of the dielectric layer was 13 μm. The UV cutoff edges gradually shift towards longer wavelength with increasing the preparation temperature of glass powders and the firing temperature of dielectric layers.  相似文献   

19.
J.W Park  Haydn Chen 《Journal of Non》1980,40(1-3):515-525
The infrared absorption spectra of sodium-disilicate glasses containing various amounts of Fe2O3 ([Na2O · 2 SiO2]1−x [Fe2O3]x, where X = 0.05, 0.1 and 0.2) were investigated in the wavenumber range from 200–2000 cm−1. The addition of Fe2O3 to the sodium-disilicate glass does not seem to introduce any new absorption band as compared with the spectrum of a pure sodium-disilicate glass; nevertheless, a general shift of the existing absorption bands toward lower wavenumbers is observed. The amount of shift is, in fact, proportional to the content of Fe2O3 in the glass. This observation is consistent with the recently proposed structural model for the bonding of Fe3+ ions in the iron-sodium-silicate glass system.

Annealing of 20 mol% iron oxide glasses at 550 and 580°C produced an extra sharp infrared absorption peak at about 610 cm−1 wavenumber. This new peak is believed to be related to the crystallized particles of the glass as concluded from both a scanning electron micrograph and an electron diffraction pattern.  相似文献   


20.
Melts with the basic compositions 10Na2O · 10MgO · xAl2O3 · (80−x)SiO2 (x=0, 5, 10, 15 and 20), 10Na2O · xMgO · 10Al2O3 · (80−x)SiO2 (x=5, 10, 15 and 20) and xNa2O · 10MgO · 10Al2O3 · (80−x)SiO2 (x=5, 10 and 15) all doped with 0.25 mol% Fe2O3 were studied using square-wave voltammetry. The temperatures applied were in the range of 1000–1600 °C. The square-wave voltammograms recorded show peaks caused by the reduction of Fe3+ to Fe2+. The attributed peak potentials measured decreased linearly with decreasing temperatures. Increasing the MgO-concentration led to more negative peak potentials. Introducing alumina in the melt first resulted in less negative peak potentials. If the molar Al2O3-concentration is equal to that of Na2O (=10 mol%) the peak potentials are least negative. Further increase of the Al2O3-concentration led to more negative peak potentials. The variation of the Na2O-concentration led to a maximum in the peak potentials at an Na2O-concentration of 10 mol%. An empirical formula which allows the calculation of standard potentials from the chemical composition is proposed. Furthermore, a structural explanation for the effect of the chemical composition is given. Especially, the incorporation of Al2O3 as AlO4-tetrahedra at [Al2O3] < [Na2O] and as network modifier at larger concentrations was structurally explained by the similarities of Fe2+ and Mg2+, with respect to cation radii and metal–oxygen bond lengths.  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