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 296 毫秒
1.
MgAl2O4 spinel precursor was prepared using a heterogeneous sol–gel process. The effect of high-energy milling on the precursor decomposition and spinel formation was investigated. The milling decreased the Al(OH)3 dehydroxylation temperature from 190° to about 130°C. The activation energy for spinel formation decreased from 688 kJ/mol for the as-prepared precursors to 468 kJ/mol for the precursors milled for 5 h. Milling of the precursor lowered the incipient temperature of spinel formation from 900° to 800°C, and the temperature of complete MgAl2O4 spinel formation from >1280° to ∼900°C.  相似文献   

2.
Calcium silicate hydrate (C-S-H) can be viewed as a solid solution, 0.833Ca(OH)2.SiO2.0.917H2O-xCa(OH)2, at equilibrium at 30°C. On this basis, the change in Gibbsfree energy (ΔGr) in the solid-solution reaction was calculated from solubility duta for C-S-H in water. The change in ΔGr with real ratio decreased notably for the higher calcium contents (CaO/Si021.7; ×0.867). Thermochemical values for C-S-H (CaO/SiO2=1.7) were estimated to be ΔH°=-2890 kJ/mol, ΔG°=-2630 kJ/mol, and S°=200 J1/mol.K at 298 K .  相似文献   

3.
A coating approach for synthesizing 0.9Pb(Mg1/3Nb2/3)O3–0.1PbTiO3 (0.9PMN–0.1PT) and PMN using a single calcination step was demonstrated. The pyrochlore phase was prevented by coating Mg(OH)2 on Nb2O5 particles. Coating of Mg(OH)2 on Nb2O5 was done by precipitating Mg(OH)2 in an aqueous Nb2O5 suspension at pH 10. The coating was confirmed using optical micrographs and zeta-potential measurements. A single calcination treatment of the Mg(OH)2-coated Nb2O5 particles mixed with appropriate amounts of PbO and PbTiO3 powders at 900°C for 2 h produced pyrochlore-free perovskite 0.9PMN–0.1PT and PMN powders. The elimination of the pyrochlore phase was attributed to the separation of PbO and Nb2O5 by the Mg(OH)2 coating. The Mg(OH)2 coating on the Nb2O5 improved the mixing of Mg(OH)2 and Nb2O5 and decreased the temperature for complete columbite conversion to ∼850°C. The pyrochlore-free perovskite 0.9PMN–0.1PT powders were sintered to 97% density at 1150°C. The sintered 0.9PMN–0.1PT ceramics exhibited a dielectric constant maximum of ∼24 660 at 45°C at a frequency of 1 kHz.  相似文献   

4.
Thermodynamic values for PUO1.5 were assessed using an improved method for estimating fef ° 1.5 and new data for S°298 1.5. Based on the assessment, a value of ΔH°298, 1.5=–828 kJ/mol is recommended. Measurements of (CO) pressure over the nominal equilibrium 1.5+ 1.5+ C were performed between 1348 and 1923 K, yielding pressures between 0.644 and 11600 Pa. Second- and third-law analyses were used to obtain a value for ΔH°298 1.5=–93.3°3.3 kJ/mol.  相似文献   

5.
Oxygen Setf-diffusion coefficients have been measured in single crystals of N2O3 doped with Mg or Ti under an oxygen partial pressure of 20 kPa in the temperature range 1400° to 1700°C. Diffusion coefficients in Mg-doped crystals obey the equation Do (cm2/s) = 1 × 1011∼ (1×1013) exp[−915 ± 50(kJ/mol)/ RT ]. The diffusivity of oxygen in Ti-doped A12O3 is lower than Mg-doped A12O3. A vacancy mechanism explains these results.  相似文献   

6.
AlN, Al2OC, and the 2 H form of SiC are isostructural. Both SiC–AlN and AlN–Al2OC form homogeneous solid solutions above 2000° and 1950°C, respectively. The kinetics of phase separation in the two systems, however, are quite different. Interdiffusion in both SiC–AlN and AlN-Al2OC systems was examined in the solid-solution regime in an attempt to elucidate differences in the kinetics of phase separation that occur in the two systems when annealed at lower temperatures. Diffusion couples of (SiC)0.3(AlN)0.7/(SiC)0.7(AlN)0.3 and (AlN)0.7(Al2OC)0.3/(AlN)0.3(Al2OC)0.7 were fabricated by hot pressing and were annealed at high temperatures by encapsulating them in sealed SiC crucibles to suppress loss due to evaporation. Interdiffusion coefficients in (SiC)0.3-(AlN)0.7/(SiC)0.7(AlN)0.3 diffusion couples were measured at 2373, 2473, and 2573 K, and the corresponding activation energy was determined to be 632 kJ/mol. (AlN)0.7(Al2OC)0.3/ (AlN)0.3(Al2OC)0.7 samples were annealed at 2273 K. The interdiffusion coefficient measured in the AlN–Al2OC system was much larger than that in the SiC–AlN system.  相似文献   

7.
The oxidation process of Si2N2O, prepared by a hot isostatic pressing technique, has been studied by the thermogravimetric method. The oxidation has been performed in oxygen for 20 h in the temperature range 1300° to 1600°C, producing oxide scales of amorphous SiO2 and α-cristobalite. The weight gain for T 1350°C does not begin to follow a parabolic rate law, until a certain time, t 0. The A 0 parameter in the parabolic rate law, (Δ w / A 0)2= K p t + B , represents the cross section area, A , through which the oxygen diffuses; in the derivation of this law A is assumed to be constant during the experiment. If crystallization occurs during the oxidation process, A will decrease with time. A function, A ( t ), describing the time dependence, has been developed and incorporated into the parabolic rate law, yielding a new rate law, which reads Δ W/A 0= a arctan √ bt + c √ t . This new rate law is valid in the time interval t < t 0, whereas, for t > t 0, the oxidation process follows the equation (Δ w/A 0)2= K °p t + B 0. The relation of the latter equation to the common parabolic rate law is described. All of the oxidation curves are described by these equations. The activation energy of the oxygen diffusion (and of the oxidation ( K p)) is found to be 245 ± 25 kJ/mol, which is consistent with literature values reported for oxygen diffusion.  相似文献   

8.
A new route for preparing hydroxyapatite (Ca10(PO4)6(OH)2) bioceramic has been described. An amorphous, nanosized, and carbonate-containing calcium phosphate powder that had a Ca:P ratio of 1.67 was synthesized from calcium diethoxide and phosphoric acid in ethanol via a sol-gel method. The powder was pressed at 98 MPa into green specimens and then heated to a temperature range of 500°-1300°C. At 600°C, the powder crystallized to a carbonated hydroxyapatite and a trace of ß-tricalcium phosphate before converting to hydroxyapatite at 900°C. The thermal crystallization was associated with grain growth, shrinkage, and active surface diffusion. The activation energy of grain growth was 37 ± 2 kJ/mol. After sintering at 1100°C, the decomposition of carbonated hydroxyapatite generated a microporous ceramic with an average pore size of 0.2 µm and an open porosity of 15.5%. This microporous bioceramic can be used as a bone filler.  相似文献   

9.
Kinetics of Barium Titanate Synthesis   总被引:7,自引:0,他引:7  
Reaction curves were obtained at various temperatures and concentrations for the formation of BaTiO3 from particulate titania in Ba(OH)2 solution. Kinetic analyses were performed by constructing mathematical models which took into account the particle size distribution of the reactant titania for both the topochemically-rate-controlled and the diffusion-rate-controlled reactions. At [Ba(OH)2] > ca. 0.1 M the rate-controlling step is the Ba reaction with TiO2 at the interface. The measured activation energy is 105.5 kJ/mol. The rates are independent of Ba(OH)2 concentration, indicating that the TiO2 interface is saturated. At [Ba(OH)2] < ca. 0.1 M the rate-determining step shifts to diffusion through the product BaTiO3 layer, the rates are concentration dependent, and the BaTiO3 particle sizes are inversely proportional to the Ba(OH)2 concentrations used.  相似文献   

10.
Isothermal transformation kinetics and coarsening rates were studied in unseeded and alpha-Al2O3-seeded γ-Al2O3 powders heated in dry air and water vapor. Unseeded samples heated in dry air transformed to alpha-Al2O3 with an activation energy of 567 kJ/mol. Seeding with alpha-Al2O3 increased the transformation rates and reduced incubation times by providing low-energy sites for nucleation/growth of the alpha-Al2O3 transformation. The activation energy for the transformation was reduced to 350 kJ/mol in seeded samples heated in dry air. Seeded samples completely transformed to alpha-Al2O3 after 1 h at 1050°C when heated in dry air compared to 1 h at 925°C when heated in saturated water vapor. The combined effects of a lower nucleation barrier due to seeding and the increased diffusion due to water vapor reduced the activation energy for the transformation by 390 kJ/mol and the transformation temperature by ∼225°C compared to the unseeded samples heated in dry air. The accelerated kinetics is believed to be due to increased surface diffusion.  相似文献   

11.
Enthalpy of Formation of Zircon   总被引:1,自引:0,他引:1  
Using high-temperature solution calorimetry in molten 2PbO. B2O3, the enthalpy of reaction of the formation of zircon, ZrSiO4, from its constituent oxides has been determined: Δr H 977(ZrSiO4) =−27.9 (± 1.9) kJ/mol. With previously reported data for the heat contents of ZrO2, SiO2, and ZrSiO4 and standard-state enthalpies of formation of ZrO2 and SiO2, we obtain Δf H °298· (ZrSiO4) =−2034.2 (±3.1) kJ/mol and Δf G °298 (ZrSiO4) =−1919.8 kJ/ mol. The free energy value is in excellent agreement with a range previously estimated from solid-state reaction equilibria. At higher temperature also the data are in close agreement with existing data, though the data sets diverge somewhat with increasing T . The limitations of the data for predicting the breakdown temperature of zircon into its constituent oxides are discussed.  相似文献   

12.
Calcium hexa-aluminate (CaO·6Al2O3) has been prepared from calcium nitrate and aluminum sulfate solutions in the temperature range of 1000°–1400°C. A 0.3 mol/L solution of aluminum sulfate was prepared, and calcium nitrate was dissolved in it in a ratio that produced 6 mol of Al2(SO4)3·16H2O for each mole of Ca(NO3)2·4H2O. It was dried over a hot magnetic stirrer at ∼70°C and fired at 1000°–1400°C for 30–360 min. The phases formed were determined by XRD. It was observed that CaO·Al2O3 and CaO·2Al2O3 were also formed as reaction intermediates in the reaction mix of CaO·6Al2O3. The kinetics of the formation of CaO·6Al2O3 have been studied using the phase-boundary-controlled equation 1 − (1 − x )1/3= K log t and the Arrhenius plot. The activation energy for the low-temperature synthesis of CaO·6Al2O3 was 40 kJ/mol.  相似文献   

13.
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   

14.
A mechanistic model for the kinetics of hydrolysis of α-tricalcium phosphate (α–Ca3(PO4)2 or α-TCP) to hydroxyapatite (Ca10− x (HPO4) x (PO4)6− x (OH)2− x or HAp) has been developed. The model is based on experimental hydrolysis rate data obtained using isothermal calorimetry. Analysis of the kinetic data according to the general kinetics models in terms of the fractional degree of reaction and time suggests the hydrolysis to be controlled by different rate-limiting mechanisms as reaction proceeds. Initially, the hydrolysis kinetics depend on the surface area of the anhydrous α-TCP. Subsequently, they change to a dependence on the rate of HAp product formation controlled by a nucleation and growth mechanism. The model predicts that HAp nuclei form at essentially one time and growth occurs in two dimensions, leading to a platelike morphology. The change in the reaction mechanism occurs at a fractional degree of hydrolysis, which does not change significantly with temperature in the range of 37°–56°C.  相似文献   

15.
Single crystals of oilivine, (Mg0.9Fe0.1)2SiO4, have been oxidized in air at temperatures between 700° and 1100°C for times from 0.5 to 100 h. Both an internal and an external oxidation layer developed. Transmission and analytical electron microscopy observations reveal that the internal oxidation layer is composed of precipitates of magnetite plus amorphous silica, which nucleated heterogeneously on dislocations and grew in an Fedepleted matrix of olivine. Rutherford backscattering spec-trometry (RBS) demonstrates that the thin external oxidation layer is free of Si; that is, it is made up of Mg-Fe oxide phases. Thus, the oxidation process is primarily controlled by diffusion of Fe2+ and Mg2+ ions toward the surface with Si4+ and O2- remaining largely immobile. The kinetics of oxidation, as determined from RBS analyses of the external oxidation layer, are parabolic with an activation energy of 140 kJ/mol. Although this activation energy is lower than that reported for self-diffusion of Mg in Mg2SiO4, the diffusivity calculated from the reaction rate constant is in good agreement with published values for lattice diffusion of Mg in the limited temperature range in which data overlap. However, the rate of accumulation of Fe in the external layer is more rapid than expected for lattice diffusion, indicating that the transport of Fe is dominated by short-circuit diffusion along the precipitate complexes which decorate dislocations.  相似文献   

16.
Thermal reactions of mixtures of ultrafine particles of magnesium hydroxide (Mg(OH)2) and kaolinite in a composition of MgO:Al2O3:2SiO2 were investigated to obtain dense cordierite ceramics at temperatures <1000°C. While heating the mixture of kaolinite and Mg(OH)2 with the equivalent of 2 mass% of boron oxide (B2O3) (in the form of magnesium borate, 2MgOB2O3), an amorphous phase formed at a temperature of ∼850°C after thermal decomposition. Firing the mixture at a temperature of 900°C yielded dense ceramics with an apparent porosity of almost zero. The addition of B2O3 promoted the densification at 850°-900°C and accelerated the crystallization of alpha-cordierite. The specimen with 3 mass% of B2O3 that was fired at a temperature of 950°C showed a linear thermal expansion coefficient of ∼3 × 10−6 K−1, a bending strength of >200 MPa, and a relative dielectric constant of 5.5 at 1 MHz. These cordierite ceramics may be used as substrate materials for semiconductor interconnection applications.  相似文献   

17.
The growth behavior, time of nucleation, and morphology of Ca(OH)2 crystals formed during the hydration of Ca3SiO5, at 15°, 25°, and 35°C at water-solid ratios ( w/s ) from 0.3 to 5.0 were studied by optical microscopy. In samples with w/s >0.5 growth of Ca(OH)2 in the c -axis direction is initially dominant. Growth in this direction ends after a few hours, but growth perpendicular to the c axis continues for several days and produces a dendritic morphology. Growth behavior is not so well defined for w/s <0.5, in part because of the large number of unhydrated particles engulfed. Increasing temperature resulted in an increase in the number of Ca(OH)2 nuclei and a decrease in nucleation time and crystal size. Increasing the w/s ratio improved the euhedral character of the Ca(OH)2 crystals, decreased the number of engulfed Ca3SiO5 particles, and increased the nucleation time. Dendritic morphology was most pronounced in the samples for which w/s = 1. Growth rates and the ultimate size of the Ca(OH)2 crystals varied within a given sample. The effects of temperature and the w/s ratio on the heat evolved during the hydration were studied by isothermal calorimetry. The times of nucleation of crystalline Ca(OH)2 estimated from calorimetry were similar to those derived from growth curves determined by optical microscopy.  相似文献   

18.
The liquidus-solidus relations along the join Ca2SiO4-Ca(OH), in the system CaO-SiO2-H2O have been determined at 1000 atm up to 1110°C. This join is binary and contains the calcium silicate hydrate, calciochondrodite, Ca5-(SiO4(OH)2. Calciochondrodite melts incongruently to Ca2SiO2+ liquid (composition 23 wt% Ca2Si04) at 955°C. The eutectic between calcium hydroxide and calciochondrodite lies at 13% Ca2Si04 and 822°C. Preliminary experiments, also at 1000 atm, in the ternary system CaO-Ca2Si04-Ca(OH), indicate that the eutectic at which the fields of primary Ca(OH)2, CaO, and Ca2(Si04)2(OH)2 meet is close to the CaO-Ca. (OH), side of the triangle at approximately 805° C. The ternary reaction point Ca2SiOl+ liquid ⇌Ca5(SiO4)2(OH)2+ CaO + liquid is believed to lie in the low-CaO (<5%) high-Ca(OH)2 (>70%) part of the system.  相似文献   

19.
The formation of crystalline Li4SiO4 and Li2SiO3 from lithium orthosilicate glasses has been studied by means of in situ high-temperature X-ray diffraction. The first phase that crystallizes from the glass could not be identified, but was followed by the transformation to crystalline orthosilicate and metasilicate. Orthosilicate formation was tracked at temperatures between 600° and 650°C, whereas at higher temperatures, the formation of a small amount of crystalline metasilicate occurs. The kinetics of the initial phase of both crystallization processes are described by the Avrami–Erofeev equation, resulting in activation energies of 90 kJ/mol for the formation of the unidentified phase, and 68 kJ/mol for the formation of Li4SiO4. The rate constant for the crystallization of the unidentified phase is 0.014 s−1 at 510°C, equaling that of the orthosilicate formation at 630°C. After isothermal heat treatment for 100–800 s, depending on temperature, 80%–95% of the sample is crystallized and further crystallization is controlled by diffusion in both cases.  相似文献   

20.
Grain growth in a high-purity ZnO and for the same ZnO with Bi2O3 additions from 0.5 to 4 wt% was studied for sintering from 900° to 1400°C in air. The results are discussed and compared with previous studies in terms of the phenomenological kinetic grain growth expression: G n— G n0= K 0 t exp(— Q/RT ). For the pure ZnO, the grain growth exponent or n value was observed to be 3 while the apparent activation energy was 224 ± 16 kJ/mol. These parameters substantiate the Gupta and Coble conclusion of a Zn2+ lattice diffusion mechanism. Additions of Bi2O3 to promote liquidphase sintering increased the ZnO grain size and the grain growth exponent to about 5, but reduced the apparent activation energy to about 150 kJ/mol, independent of Bi2O3 content. The preexponential term K 0 was also independent of Bi2O3 content. It is concluded that the grain growth of ZnO in liquid-phase-sintered ZnO-Bi2O3 ceramics is controlled by the phase boundary reaction of the solid ZnO grains and the Bi2O3-rich liquid phase.  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