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 15 毫秒
1.
A comparative study of infrared absorption due to H2O and D2O impurities in a fluorozirconate glass (53ZrF4·20BaF2·4LaF3·3AlF3·20NaF) was carried out. The H2O and D2O were introduced into the glass by reaction of the surface at 260°C with H2O and D2O vapor entrained in a stream of N2. Reaction with H2O produced IR absorption bands at 2.9 μm (O–H stretch) and 6.1 μm (H2O bend). Reaction with D2O produced bands at 3.9 μm (O–D stretch) and 8.3 μm (D2O bend). The ratios of the corresponding D2O/H2O peak frequencies are 0.74 for both the stretching and bending vibrations, in good agreement with the value of 0.727 predicted from the difference in the OH and OD reduced masses.  相似文献   

2.
Na2O–3SiO2 glasses with up to ≅12 wt% water were prepared under high-pressure, hydrothermal conditions and their electrical conductivities were measured. The conductivity (σ) was found to depend on H2O content in a manner similar to the "mixed-alkali" effect. At constant temperature, σ decreased initially with increasing H2O content to a minimum at 3≅4 wt% H2O and increased with further increase in water content. Infrared spectroanalysis of these glasses was made to determine the type of water present in the glass and the results were interpreted in the context of the measured activation energy and preexponential factor of dc conductivity. The sodium ion diffusion coefficient for a glass containing 0.76 wt% H2O was less than that of water-free glass and in agreement with the electrical conductivity results.  相似文献   

3.
Examination of tobermorite, 4–SCaO.5SiO2.-5H2O, and xonotlite, 5CaO.5SiO2.H2O, by infrared absorption revealed striking structural similarities and differences in the two phases. In the 8- to 15- μ region, the absorption was the same for both; the differences arose from the manner in which water and hydroxyl ions were bonded. Tobermorite exhibited strong absorption at 6.2 μ , a band which is generally associated with interlayer water, and at 2.9 μ , a band generally attributed to bonded (OH). The mineral xonotlite did not show these two bands but contained the band at 2.75 μ generally associated with free (OH). Synthetic xonotlite prepared at 300°C. was essentially the same as the mineral, but samples prepared at progressively lower temperatures exhibited the 2.9- μ band in increasing intensity. The fibrous form of tobermorite showed a band at 6.5 to 7.0 μ which increased in intensity with increasing amount of CaO in the solid; this band was also found in the 14-a.u. 1.0 CaO: SiO2 hydrate, but not in xonotlite. The great volume stability of xonotlite during drying and wetting is readily explained on the basis of the present results. Shrinkage of tobermorite during drying at temperatures up to 650°C. may be due to removal of both interlayer water and bonded (OH). The changes in absorption during drying at room temperature were too small, however, to permit drawing any conclusions. Similarities and differences between tobermorite and certain clays are discussed.  相似文献   

4.
The reaction of rare-earth (RE; Y, Er, and Yb) chloride hydrates in 1,4-butanediol at 300°C for 2 h gave mixtures of RE(OH)2Cl and RE2O3· x H2O, and the products were composed of irregularly shaped particles. A prolonged reaction (10 h) yielded a mixture of RE(OH)2Cl and RE2O3· x H2O for Er or Y, but phase-pure RE2O3· x H2O was obtained for Yb. The product for Yb comprised needle-shaped single crystals of Yb2O3· x H2O with a width of 0.2–0.6 μm and a length of 5–15 μm. The Yb2O3· x H2O phase decomposed to Yb2O3 at 350°–500°C, preserving the needle-shaped morphology; this was maintained even after calcination at 1100°C. Single crystals of Yb2O3 obtained by the calcination of Yb2O3· x H2O at 500°C had very small voids and the voids were enlarged to 35 Å in diameter by calcination at 800°C.  相似文献   

5.
The phase relations were established experimentally for the system CaO-Al2O3-P2O5-H2O at 200°C and 1710 kPa. The quaternary compound, crandallite, CaAl3(PO4)2(OH)5· H2O, was found to be stable. Compatibility joins in the system were determined. The phase relations are presented on the isothermal-isobaric 90 wt% water plane and by projecting the primary fields of the liquidus surface onto the same plane.  相似文献   

6.
Phase equilibria have been determined in the system CaO-Al2O3-H2O in the temperature range 100° to 1000°C. under water pressures of up to 3000 atmospheres. Only three hydrated phases are formed stably in the system: Ca(OH)2, 3CaO·Al2O3·6H2O, and 4CaO·3Al2O3-3H2O. Pressure-temperature curves delineating the equilibrium decomposition of each of these phases have been determined, and some ther-mochemical data have been deduced therefrom. It has been established that both the compounds CaO·Al2O3 and 3CaO·Al2O3 have a minimum temperature of stability which is above 1000°C. The relevance of the new data to some aspects of cement chemistry is discussed.  相似文献   

7.
Dehydration of Hydrous Zirconia with Methanol   总被引:13,自引:0,他引:13  
The washing of hydrous zirconia with alcohols to reduce the incidence of hard agglomerates on subsequent drying is well known. The results of methanol dehydration of hydrous zirconia (zirconium hydroxide), [Zr4(μ-OH)8(OH)8(H2O)8]˙ xH2O, show that only μ-OH groups are unaffected. This suggests two things: First, the removal of nonbridging hydrooxo groups and water with alcohols such as methanol leads to a reduction/elimination of hard agglomerates. Second, hard agglomerate formation is associated with condensation reactions involving nonbridging hydroxo groups (Zr-OH+HO-Zr→Zr-O-Zr+H2O).  相似文献   

8.
In the modified chemical vapor deposition (MCVD) process for making glass fibers, most of the hydrogen coming into the reaction from hydrogen-bearing impurities in the starting materials is not incorporated in the glass. It is instead mostly converted to HCl, which is not absorbed by the newly formed silica particles, and passes out the exhaust stack. The residual partial pressure of H2O formed in equilibrium with HCl in the presence of O2 and Cl2 accounts quantitatively for the OH appearing in the glass when the known solubility of H2O in silica is considered. The H2O/HCl equilibrium is quenched at a temperature below that of the reaction zone at a value for which the rate of reaction approximates the transit time across the deposition zone. Quantitative agreement is obtained for published OH concentrations produced by SiHCl3 doping.  相似文献   

9.
The precipitation process of solid phases Mg3(OH)5CI-4H2O (phase 5), Mg2(OH)3CI-4H2O (phase 3), and Mg(OH)2 was followed by the addition of NaOH water solution in MgCl2 water solutions of different concentrations (0.001 to 4.8 mol dm−3) and characterized by chemical, potentiometric, coulometric, and X-ray diffraction analyses. The concentration range in which the precipitation of solid phases occurs was determined. The phase distributions relative to the pH of solution and concentrations of magnesium and chloride were defined by the equilibrium diagram. The approximate solubility products of stable solid phases formed at different ionic strengths and at 293 K were determined.  相似文献   

10.
Phase equi ibria in the system MgO-MgC 2-H2O at 2°±3°C were determined and the reactions by which the equi ibrium phases, 5Mg(OH)2 MgC 2 8H2O and 3Mg(OH)2 MgC 28H2O, deve op were studied by X-ray diffractometry. As reactive MgO is disso ved by magnesium ch oride so utions, a thixotropic suspension is converted to a ge, which then crysta izes to form the ternary oxych oride phases. Insufficient y active MgO disso ves more s ow y so that, in an open system, bu k composition can shift by evaporation of water, resu ting in crysta ization of a nonequi ibrium phase assemb age, with residua magnesium ch oride so ution and unreacted MgO. Imp ications of these nonequi ibrium reactions for the performance of magnesium oxych oride cements are discussed.  相似文献   

11.
Factors influencing the low-temperature formation of AIPO4 and its precursor phases, AIPO4· x H2O (1 x 2), were investigated. AIPO4 formed by reaction between 33.3 wt% H3PO4 solution and alumina. Five aluminas (three anhydrous and two hydrated) were utilized. Each differed in particle size, surface area, and crystallinity. The reaction temperatures investigated were 113°, 123°, and 133°C. The high-surface-area aluminas were sufficiently reactive in the phosphoric acid solution at these temperatures to produce crystalline reaction products. However, only hydrated forms of AIPO4, AIPO4· x H2O (1 x 2), crystallized directly out of solution. x generally decreased as the curing temperature was increased. Upon dehydration of these hydrated reaction products, anhydrous AIPO4 was formed, primarily in the berlinite and/or cristobalite modifications. Both the temperature of reaction and the alumina used influence the hydrates that form. In turn, the hydrates which form, the macroscopic assemblages into which they may crystallize, and the morphologies of the crystallites all affect the polymorphic form and the crystallinity of the anhydrous AIPO4 phase ultimately produced on dehydration. Phase-pure and highly crystalline AIPO4-cristobalite (the high-temperature modification) was formed by the dehydration of AIPO4·H2O at a temperature as low as 113°C.  相似文献   

12.
Values of the spectral absorption coefficient (α) of liquid aluminum oxide were determined by transmission of a pulsed dye laser beam incident on continuous-wave (CW) CO2-laser-melted pendant drops attached to sapphire filaments. Measurements were made on molten drops of Verneuil sapphire at wavelengths of 0.450 and 0.633 μm, at ambient oxygen partial pressures from 10-10 to 1 bar in eight pure gases (Ar, CO, CO2, H2, H2O, HCI, N2 and O2), in CO/CO2 mixtures, and in H2/H2O mixtures, and at a temperature of ca. 2400 K. Specimens contaminated with iron, magnesium, silicon, and tungsten were also investigated in an oxygen atmosphere. At a wavelength of 0.633 μm, the value of α was greater than 50 cm-1 under reducing or inert gas conditions. It decreased to a minimum at intermediate oxygen partial pressures of 5 × 10-5 bar in CO/CO2 mixtures and 5 × 10-3 bar in H2/H2O mixtures, and increased at larger oxygen partial pressures. The specimens were opaque (α > 55 cm-1) in hydrogen, in HCI at pressures above 0.04 bar. Specimens contaminated with 5000-10000 ppm of Fe, Mg, Si, or W were also opaque. At a wavelength of 0.45 μm the liquid aluminum oxide specimens were opaque in Ar and oxygen, and gave α= 46 cm-1 in CO2. The dynamic response when the ambient gas was changed from CO2 to argon showed that the transmission maximum for = 0.45 μm was at p (O2) < 0.1 bar.  相似文献   

13.
The hydration of two high replacement composite cements (3:1 blast furnace slag:ordinary Portland cement (BFS:OPC), and 3:1 pulverized fuel ash:OPC (PFA:OPC)) with the addition of both SnCl2 and SnCl4 has been investigated and the results from X-ray diffraction (XRD) and scanning electron microscopy (SEM) with energy-dispersive spectroscopy (EDS) are presented. Adding 5% or 1% SnCl2·2H2O or SnCl4·5H2O to the mix water resulted in the formation of Friedel's salt, Ca3Al2O6.CaCl2·10H2O, and calcium hydroxo-stannate CaSn(OH)6, which also involved the consumption of calcium hydroxide. After 90 days hydration at lower levels of addition (i.e., 1%) there was no longer evidence for CaSn(OH)6, indicating that it too had been consumed in the pozzolanic reaction due to the lack of calcium hydroxide present. Results from SEM and EDS showed that bright regions between the BFS or PFA grains were tin containing and they were incorporated into the hydrated cement matrix. The tin was, therefore, localized rather than spread throughout and intimately incorporated into the microstructure.  相似文献   

14.
Attenuated total reflectance Fouriertransform infrared (ATR-FTIR) spectra were measured in the region from 4300 to 400 cm−1 for a hydrated Na2O–SiO2 glass containing 35 wt% water. The Si–OH bending vibration mode was observed. It was found that the incorporated water, molecular water as well as hydroxyls, affected the Si–O vibrations. The effect of incorporated water upon the glass structure is discussed.  相似文献   

15.
Solubility in the fully hydrated CaO–SiO2–H2O system can be best described using two ideal C-S-H-(I) and C-S-H-(II) binary solid solution phases. The most recent structural ideas about the C-S-H gel permit one to write stoichiometries of polymerized C-S-H-(II) end-members as hydrated precursors of the stable tobermorite and jennite minerals in the form of 5Ca(OH)2·6SiO2·5H2O and 10Ca(OH)2·6SiO2·6H2O, respectively. For thermodynamic modeling purposes, it is more convenient to express the number of basic silica and portlandite units in these stoichiometries using the coefficients n Si and n Ca. Thermodynamic solid-solution aqueous-solution equilibrium modeling by applying the Gibbs energy minimization (GEM) approach shows the best generic fits to the available experimental solubility data at solid 0.8 < Ca/Si < 2.0 if both stoichiometry and thermodynamic constants of the end-members are normalized to n Si= 1.0 ± 0.3. Recommended stoichiometries and thermodynamic data for the C-S-H end-members provide a reliable basis for the subsequent multicomponent extension of the ideal C-S-H solid solution model by incorporation of end-members for the (radio)toxic elements or trace metals.  相似文献   

16.
Internal friction was measured for NaPO3 glasses containing from 0.0004 to 0.146 wt% H2O. In glasses containing ≤ 0.001 wt% H2O, the only internal friction peak observed was that resulting from the Na+ ions, i.e. the alkali peak; it was ∼ 30% larger than in as-melted glasses. The second peak normally observed in as-melted NaPO3 glass was completely absent when the water content was < 0.001 wt%, consistent with the interpretation that it is caused by a reorientation of proton-Na+ elastic dipoles. The concept that the second peak in as-melted alkali phosphate and silicate glasses results from nonbridging oxygen ions is not substantiated by these studies of NaPO3 glasses, which contain a high percentage of these ions but do not exhibit this peak when the proton concentration is sufficiently low.  相似文献   

17.
The compound compositions of four aluminous cements were determined on anhydrous as well as hydrated specimens which had been heat-treated at temperatures between room temperature and 1400° C. Phases were identified by X-ray diffraction and differential thermal analysis. Specimens were also tested for transverse strength, dynamic modulus of elasticity, and thermal length change. A study of the dehydration characteristics of CaO - Al2O8 - 10H2O3 3CaO.Al2O3. 6H2O, and Al2O3. 3H2O was included. The data indicated that CaO. Al2O3 10H2O was the primary crystalline hydrate formed in the cements at room temperature. At 50° C., 3 CaO Al2O3-6H2O and Al2O3. 3H2O were formed as by-products of the dehydration of CaO.Al2O3.10H2O. When heated alone in an open system, CaO.Al2O3.10H2O did not convert to 3CaO. Al2O3. 6H2O and A12O3. 3H2O. A correlation between the mechanical properties and compound compositions was noted.  相似文献   

18.
Rutile or anatase may be depolymerized and complexed by sequential treatment with (i) H2SO4/(NH4)2SO4, (ii) H2O, and (iii) catechol/NH4OH to produce the intermediate (NH4)2(Ti(catecholate)3) · 2H2O. Treatment with Ba(OH)2· 8H2O leads to an acid-base reaction generating Ba(Ti(catecholate)3) · 3H2O, in which the Ba:Ti ratio is held at 1:1 at the molecular level. Calcination produces BaTiO3 powder.  相似文献   

19.
The nature of the low-temperature inversions γ-α' and α'-β was investigated by various techniques: hydrothermal and "dry" quenching runs, differential thermal analysis at atmospheric and elevated nitrogen pressures, X-ray diffractometer patterns obtained at elevated temperatures, "static" pressure techniques, and infrared absorption spectrometry. A revised energy-temperature diagram is presented for Ca2SiO4, with the transition γ' to α' taking place at about 725°C. and the α'-β transition, although not reversible at an exact temperature, taking place at about 670° C. At low water pressures (2000 lb. per sq. in.) the inversion γ-α' was placed at 675°C. Attempts to extrapolate the value obtained at 2000 lb. per sq. in. to obtain a more accurate reversible inversion temperature at atmospheric pressure, although limited in accuracy by the reliability of heat-of-transition data, would indicate a temperature of about 725° C. at atmospheric pressure. Three new compounds, 8CaO.3SiO2 -3H2O (X), 6CaO 3SiO2.H2O (Y), and 9CaO-6SiO2 H2O (Z), were found to be stable above 700°C. at H2O pressures greater than 7500 lb. per sq. in.  相似文献   

20.
Single-crystal X-ray and electron-diffraction studies show the existence in one polymorph of 4CaO.Al2O3. 13H2O of a hexagonal structural element with α= 5.74 a.u., c = 7.92 a. u. and atomic contents Ca2(OH)7- 3H2O. These structural elements are stacked in a complex way and there are probably two or more poly-types as in SiC or ZnS. Hydrocalumite is closely related to 4CaO.A12O3.13H2O, from which it is derived by substitution of CO32-for 20H-+ 3H2O once in every eight structural elements; similar substitutions explain the existence of compounds of the types 3CaO Al2O3.Ca Y 2- xH2O and 3CaO Al2O3 Ca Y xH2O. On dehydration, 4CaO.Al2O3.13H2O first loses molecular water and undergoes stacking changes and shrinkage along c. At 150° to 250°C., Ca(OH)2 and 4CaO.3Al2O3.3H2O are formed and, by 1000°C., CaO and 12CaO.7Al2O8. The dehydration of hydrocalumite follows a similar course, but no 4CaO.3Al2O3.3H2O is formed.  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