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 46 毫秒
1.
By using simple geometric concepts, general thermodynamic dependences governing the process of vitrification are derived. The thermodynamic driving forces of crystallization, ΔGf,c(T), and stabilization, ΔGst(T), melts and glasses are considered in the entire temperature interval from temperatures below the glass transition temperature, Tg, to temperatures high above the melting point, Tm. It is show that at large deviations from Tg the temperature dependence of the driving force for stabilization, ΔGst(T), is non-symmetric and expressions derived by a Taylor expansion become a poor approximation. By using an appropriate thermodynamic model, a more general expression for ΔGst(T) is obtained and the thermodynamic basis for the existence of superheated glasses is established. The thermodynamic instability of glasses below and above Tg, i.e., of supercooled and superheated glasses is discussed.  相似文献   

2.
Alkali diffusion and electrical conductivity in sodium borate glasses   总被引:2,自引:0,他引:2  
Y.H Han  N.J Kreidl  D.E Day   《Journal of Non》1979,30(3):241-252
The diffusion coefficient of sodium, 22Na, and silver, 110Ag, and electrical conductivity for sodium borate glasses (4–24 mol% Na2O) has been measured from 100°C to slightly below the glass transition temperature, Tg. Unlike silicate glasses where the self-diffusion coefficient is much larger than the impurity diffusion coefficient, DNa and DAg had close to the same magnitude in the sodium borate glasses. In some glasses, DAg was slightly larger than DNa. Na2O-Ag2O-B2O3 glasses show a relatively small mixed alkali effect, despite the significant mass difference between Ag(≈108) and Na(≈23). It is concluded that the mixed alkali effect is more dependent upon the difference in ionic radii rather than differences in mass.  相似文献   

3.
Formation and structure of titanate glasses   总被引:2,自引:0,他引:2  
Cheng Jijian  Chen Wei 《Journal of Non》1986,80(1-3):135-140
The formation of high titanium oxide (30–60 wt%) containing glasses was studied. Stable titanate glasses with high content of TiO2 and BaO could be obtained even without other glass-forming oxides. It is demonstrated that the coloration of titanate glasses with high content of TiO2 is due to oxygen loss during the melting. A systematic study of controlling melting and heat treatment conditions led to the successful decoloration of titanium oxide containing glasses. Infrared spectra and X-ray diffraction studies showed that Ti4+ in titanate glasses is in sixfold coordination. Phase separation was observed by electron microscopy when the titanate glasses were heat-treated at the temperature above Tg. The crystallization of titanate glasses is generally preceded by phase separation. The obtained crystalline phases are mainly different titanates.  相似文献   

4.
Literature data on the effect of water on the glass transition in silicate melts are gathered for a broad range of total water content cw from 3 × 10−4 to 27 wt%. In terms of a reduced glass transition temperature Tg*=Tg/TgGN, where TgGN is Tg of the melt containing cw≈0.02 wt% total water, a uniform dependence of Tg* on total water content (cw) is evident for silicate melts. Tg* decreases steadily with increasing water content, most strongly at the lowest water content where H2O is dominantly dissolved as OH. For water-rich melts, the variation of Tg* is less pronounced, but it does not vanish even at the largest water contents reported (≈27 wt%). Tg* vs. cw is fitted by a three-component model. This approach accounts for different transition temperatures of the dry glass, hydroxyl and molecular water predicting Tg* as a weighted linear combination of these temperatures. The required but mostly unknown water speciation in the glasses was estimated using IR-spectroscopy data for hydrous sodium trisilicate and rhyolite.  相似文献   

5.
A. Dahshan   《Journal of Non》2008,354(26):3034-3039
Thermal stability and crystallization kinetics of As14Ge14Se72−xSbx (where x = 3, 6, 9, 12 and 15 at.%) glasses are studied by the differential scanning calorimetry. The values of the glass transition temperature (Tg) and the peak temperature of crystallization (Tp) are found to be dependent on heating rate and antimony content. From the heating rate dependence of Tg and Tp the values of the activation energy for glass transition (Et) and the activation energy for crystallization (Ec) are evaluated and their composition dependence discussed. Crystallization studies have been made under non-isothermal conditions with the samples heated at several uniform rates. Using a recent analysis developed for non-isothermal crystallization studies, information on some aspects of the crystallization process has been obtained. The stability calculations emphasized that the thermal stability decreases with increasing the Sb content.  相似文献   

6.
Possible relationships between measures of glass stability (GS) against devitrification on heating (evaluated by the Hrubÿ parameter KH=(TchTg)/(TmTch), and the parameter Kw=(TchTg)/Tm) and a criterion of glass-forming ability (GFA) – the critical cooling rate – were investigated by computing non-isothermal crystallization for typical values of the main quantities that control crystal nucleation and growth in silicate glasses. We limit these quantities to one thermodynamic parameter – the melting entropy (ΔSm) and two kinetic parameters that control the viscosity (B and T0 in the Vogel–Fulcher–Tamman equation or Tg and in Avramov’s equation). The effect of heterogeneous nucleation and, in particular, the possible role of the surface as active substrate is tested. The results presented herein demonstrate that GS and GFA are indeed related concepts.  相似文献   

7.
Phosphate glass surfaces were nitrided by reacting them in dry ammonia at temperatures near the glass transition temperature (Tg) for up to 100 h. These treatments have significant effects on surface dependent glass properties. For example, the dissolution rate of a sodium-barium phosphate glass (Tg = 345°C) decreased by over an order of magnitude after 24 h in ammonia at Tg. X-ray photoelectron spectroscopic (XPS) analyses show that nitrogen in chemically incorporated into the glass structure at levels up to 3–5 at.%. Elemental depth profiles, obtained by XPS (for N) and by elastic recoil detection (ERD) analyses (for N and H), indicate that the surface oxynitrile layer extends to about 1 μm in depth. XPS analyses of in situ treated samples reveals the presence of several nitrogen species which affect the surface dependent properties by increasing the structural crosslink density of the glass surface.  相似文献   

8.
We show that simple expressions can be derived from the Vogel–Fulcher–Tammann (VFT) law relating the glass transition temperature Tg, the VFT temperature T0, their pressure derivatives, the steepness index of the ‘Angell plot’ and the strength parameter D of the VFT equation, in good agreement with experimental data. In the same way one can describe the dependence of the dTg/dP on the relaxation time τg chosen to define the temperature Tg. Thus, this procedure allows a consistent rescaling and comparison of pressure dependent parameters obtained from different experiments and simulations.  相似文献   

9.
R. Mathai  G. H. Frischat   《Journal of Non》1999,260(3):175-179
A glass of composition 53ZrF4–20BaF2–4LaF3–3AlF3–20NaF (Tg=260°C) was prepared by careful crucible melting. High-resolution atomic force microscopy of fracture surfaces displayed the presence of nano-pores with diameters of 20–50 nm, being 4–10 nm deep, in all glasses. It was further found that only glasses without annealing and glasses with an annealing step considerably below Tg showed a distinct pattern, i.e. ripples of ≈20 nm in diameter and an rms roughness of ≈0.6 nm. Glasses annealed either near Tg or at the temperatures of maximum nucleation or maximum crystal growth rates showed both regions with the ripple pattern and regions with nano-hillocks, growing in size with increasing annealing temperature and time. Thus these hillocks nearly reach micro-dimensions of ≈270 nm in diameter and ≈65 nm in height following a 90 min annealing step at 343°C, the temperature of maximum crystal growth. These findings give evidence that the glass system, which is thought to be one of the most suitable for fiber drawing, is much less stable against nucleation and crystallization than anticipated.  相似文献   

10.
A. Inoue  T. Zhang  T. Masumoto 《Journal of Non》1992,150(1-3):396-400
An amorphous phase with a wide supercooled liquid region, > 50 K, was found to form over wide composition ranges in the La---Al---Ni and Zr---Al---Cu systems. The largest values for the temperature span between the crystallization temperature, Tx, and the glass transition temperature, Tg, ΔTx(-TxTg), are 69 K for La55 Al25Ni20 and 88 K for Zr65Al7.5Cu27.5. The structural relaxation behavior on annealing was examined for the two amorphous alloys with the largest ΔTx values. The magnitude of the structural relaxation increases gradually with increasing annealing temperature, Ta, and then rapidly in the Ta range slightly below Tg and decreases significantly on annealing Tg. The rapid increase in the magnitude of the structural relaxation on annealing near Tg is due to the glass transition. The single-stage structural relaxation indicates that there is no distinct difference in relaxation times (atomic bonding forces) between the constituent atoms in the two metal-metal-type amorphous alloys. The existence of an optimum bonding state is thought to cause the wide supercooled liquid region for the two amorphous alloys.  相似文献   

11.
Z. P. Lu  Y. Li  S. C. Ng 《Journal of Non》2000,270(1-3):103-114
Onset temperature (solidus) Tm and offset temperature (liquidus) Tl of melting of a series of bulk glass forming alloys based on Zr, La, Mg, Pd and rare-earth elements have been measured by studying systematically the melting behaviour of these alloys using DTA or DSC. Bulk metallic glass formation has been found to be most effective at or near their eutectic points and less effective for off-eutectic alloys. Reduced glass transition temperature Trg given by Tg/Tl is found to show a stronger correlation with critical cooling rate or critical section thickness for glass formation than Trg given by Tg/Tm.  相似文献   

12.
Vanadium EXAFS spectra of 50V2O5-50P2O5 glasses with different C = V+4/Vtot have been measured. The V-O distances increase by ΔR1 = (0.03 ± 0.02) Å to ΔR2 = (0.07 ± 0.03) Å going from a glass with C = 0.64 to C = 0.84 and from C = 0.51 to C = 0.84, respectively. The EXAFS data show a basically similar structure of the vanadium sites for both the V4+ and V5+ ionic states. The density of the glasses increases with C from the value d1 = 2.81 g/cm3 for C = 0.51 to d2 = 2.92 g/cm3 for C = 0.84 indicating a more random packing of glass units.  相似文献   

13.
Measurements of solid phase dopant concentration (S) of LPCVD Si thin films as a function of substrate temperature (Ts = 500−640 ° C) and gas phase doping ratio (R = 1 × 10−5 −4 × 10−2) by SIMS indicate different behaviors of P and B in the films. A linear relation S = b(T)R is observed for B-doped film with b(T) varying from 4 to 50 depending on Ts. Boron-doped microcrystalline film has a higher doping efficiency than that of P-doped ones.  相似文献   

14.
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   

15.
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   

16.
Yuan Lirong  Yao Guoxing 《Journal of Non》1988,100(1-3):309-315
This study demonstrates that hydrolysis should be carried out in a step manner in gel synthesis. The key to the increase in the amount of water added is the control of the hydrolysis rate of Ti(OC4H9)4. The hydrolysis of Si(OC2H5)4 can be carried out at about 75°C. The amount of added water (γWI), which varied with TiO2 content (in mol%), was about 64–88% of the total amount of added water. The hydrolysis reaction should be performed at room temperature while Ti(OC4H9)4 is added. The total amount of added water (γW) is related to the amount of solvent (R). For example, if TiO2 is 40mol%, γW will vary from 3.2 to 8.0 when R varies from 0.8 to 2.0. The amount of added water was affected by the distribution of solvent in the metal alkoxides. The amount of added water can be increased when RSi(OC2H5)4 = 1, RTi(OC4H9)4 > 1. The rate of rise in temperature of the thermolysis of the dried gel should be less than 10°C per hour, and the heat treatment temperatyre is related to the TiO2 content (in mol%). Gel glasses without devitrification can only be obtained by thermolysis at 600°C from the gel with no less than 20 mol% TiO2.  相似文献   

17.
《Journal of Non》2000,270(1-3):137-146
The Ge25Ga5Se70 and Ge30Ga5Se65 pure and Pr3+-doped glasses were prepared by direct synthesis from elements and PrCl3. It was found that up to 1 mol% PrCl3 can be introduced in the Ge25Ga5Se70 and Ge30Ga5Se65 glasses. Both types of glasses with overstoichiometric and substoichiometric content of Se were homogeneous and of black color. The optical energy gap is Eoptg=2.10 eV, and the glass transition temperature is Tg=543 K for Ge25Ga5Se70 and Tg=633 K for Ge30Ga5Se65. The long-wavelength absorption edge is near 14 μm and it corresponds to multiphonon processes. Doping by Pr3+ ions creates absorption bands in transmission spectra, which can be assigned to the electron transitions from the ground 3H4 level to the higher energy levels of Pr3+ ions 3H5, 3H6, 3F2, 3F3 and 3F4, respectively. By excitation with YAG:Nd laser line (1064 nm), two intense luminescence bands (1343 and 1601 nm) were excited. The first band can be ascribed to electron transitions between 1G4 and 3H5 energy levels of Pr3+ ions. Full width at half of maximum (FWHM) of the intensity of luminescence was found to be 70 nm for (Ge25Ga5Se70)1 − x(PrCl3)x and (Ge30Ga5Se65)1 − x(PrCl3)x glasses. The FWHM in selenide glasses is lower than in halide and sulphide glasses. The second luminescence band (1601 nm) can be probably ascribed to the transitions between 3F3 and 3H4 energy levels of Pr3+ ions. The absorption and luminescence spectra of Pr3+ ions in studied glasses are slightly influenced by stoichiometry of glassy matrix. The Raman spectra of studied glasses were deconvoluted and assignment of Raman bands to individual vibration modes of basic structural units was suggested. The structure of studied glasses is mainly formed by corner-sharing and edge-sharing GeSe4 tetrahedra. The vibration modes of Ga-containing structural units were not found, they are apparently overlapping with Ge-containing structural units due to small difference between atomic weights of Ge and Ga. In the glasses with substoichiometry of Se, the Ge–Ge bonds of Ge2Se6 structural units were found. In Se-rich glasses the Se–Se vibration modes were found. In all studied glasses also ‘wrong' bonds between like atoms were found in small amounts. Maximum phonon energy of studied glasses is 320 cm−1.  相似文献   

18.
Gu Zhenan 《Journal of Non》1986,80(1-3):429-434
d-f and f-f transition bands of Pr and Sm ions in silica glasses have been studied by the absorption, excitation and emission spectra. In particular, Sm-doped silica glasses were treated with different atmospheres around the Tg temperature. The change of the valency state and the transition band shown after the treatments are discussed.  相似文献   

19.
The linewidth-broadening of the EPR spectra of Cu2+ in silicate, borate and phosphate glasses was analyzed in terms of the distribution of g| and A|| and δA|) and related to the distribution of the rigidity of the network structure. X- and K-band spectra were measured for the glasses doped with 63Cu2+ (93% abundance). The linewidth of the HFS shoulders with parallel orientation to H increased linearly with increasing m or microwave frequency. δg| and δA| showed a marked dependence on glass composition. For example, in Na2O---B2O3 glasses, on going from x (mol% of Na2O) being small through intermediate to large, δg| varied from small through large to negligibly small. In contrast to these glasses δg| was extremely large for 75PbO · 25B2O3 glass. The large δg| for the Na2O---B2O3 glassesof intermediate x was attributed to the coexistence of various borate groups competitively coordinating to Cu2+. Negligibly small δg| for 70Na2O · 30B2O3 glass and extremely large δg| for 75PbO ·25B2O3 glass, both with a narrower structural distribution, reflect regidity of the glass network. The Pb---O bonding is strong enough to distort the coordination of Cu2+-complex. The situation is the reverse in Na2O---B2O3 glasses.  相似文献   

20.
Dynamic mechanical properties of an amorphous La55Al25Ni20 alloy were measured by a forced oscillation method in the temperature range from room temperature to 453 K, just below the glass transition temperature (Tg = 481 K). The internal friction at a constant frequency 62.8 rad/s of this alloy showed a peak at about 400 K and the peak position shifts with frequency. Structural relaxation reduces the magnitude of the relaxation peak but has little affect on the peak position. Using the time-temperature superposition process, master curves for storage E′ (ln ω) and loss E″ (ln ω) moduli are constructed. Activation energy of the relaxation obtained from shift factors is low, 100 kJ/mol, which is close to that for diffusion of the Al in Al and Ni in Al. The relaxation spectra are very broad with a half width of 2 3 decades.  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