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.
The microstructure evolution and densification kinetics of alumina containing 10 and 20 vol% calcium aluminosilicate glass were studied, for sintering under vacuum and air at 1600°C. Residual porosity was always present in the air-fired samples. The kinetic analysis lent strong support to the notion that trapped gases inhibited the densification and limited the attainment of full density. The samples containing 20 vol% glass were able to reach full density during vacuum sintering. However, the samples containing 10 vol% glass contained some residual porosity even after vacuum sintering, which was attributed to the preferential volatalization of liquid phase.  相似文献   

2.
Filled glass–ceramic composites, like low-temperature co-fired ceramics (LTCC), must densify at temperatures <900°C. The densification mechanism of LTCC is often described by liquid-phase sintering. The results of this paper clearly show that densification of ceramic-filled glass–composites with a glass content above 60 wt% can be attributed to viscous sintering, which is decisively controlled by the viscosity of the glass during the heat treatment. This is demonstrated by the experimental determination of the viscosity of a MgO–Al2O3–B2O3–SiO2 glass dependent on temperature, by investigation of the wetting behavior of the glass on the ceramic filler mullite, and of the microstructural development. It was found that the glass does not wet the filler material in a temperature range up to 1000°C. Therefore, liquid-phase sintering can be excluded. Independent of any wetting effect and therefore in the absence of capillary forces, densification starts at a temperature of 750°C, which corresponds to a viscosity of 109.5 dPa·s. This densification can be attributed to viscous flow of the glass matrix composite.  相似文献   

3.
The sintering behavior of compacts of ultrafine silicon powder (0.02 to 0.1 μm particle size) was investigated. Two sintering modes occur: normal sintering associated with densification and subnormal sintering without densification. The micro-structure developed in normal sintering has a fine grain size (0.05 to 0.3 μm) and fine porosity; the grains contain stacking faults and twins. The microstructure developed in subnormal sintering exhibits larger grains (∼1 μm in size) and coarse pores. Green densities >42% of theoretical and temperatures >1100°C are required for densification. Densification follows an exponential time and temperature dependence with an activation energy of 470 kJ/mol, indicating bulk diffusion as the transport mechanism. Grain-boundary diffusion is thought to be inhibited by grain-boundary oxide films. The carbon phase-separates into discrete amorphous regions and is thought to have little effect on sintering behavior.  相似文献   

4.
Pressureless Sintering of Boron Carbide   总被引:4,自引:0,他引:4  
B4C powder compacts were sintered using a graphite dilatometer in flowing He under constant heating rates. Densification started at 1800°C. The rate of densification increased rapidly in the range 1870°–2010°C, which was attributed to direct B4C–B4C contact between particles permitted via volatilization of B2O3 particle coatings. Limited particle coarsening, attributed to the presence or evolution of the oxide coatings, occurred in the range 1870°–1950°C. In the temperature range 2010°–2140°C, densification continued at a slower rate while particles simultaneously coarsened by evaporation–condensation of B4C. Above 2140°C, rapid densification ensued, which was interpreted to be the result of the formation of a eutectic grain boundary liquid, or activated sintering facilitated by nonstoichiometric volatilization of B4C, leaving carbon behind. Rapid heating through temperature ranges in which coarsening occurred fostered increased densities. Carbon doping (3 wt%) in the form of phenolic resin resulted in more dense sintered compacts. Carbon reacted with B2O3 to form B4C and CO gas, thereby extracting the B2O3 coatings, permitting sintering to start at ∼1350°C.  相似文献   

5.
Anisotropic Grain Growth in Diphasic-Gel-Derived Titania-Doped Mullite   总被引:5,自引:0,他引:5  
Densification and anisotropic grain growth in diphasic-gel-derived, titania-doped mullite were studied. Titania enhanced initial and intermediate stage densification in diphasic mullite gels by reducing the glass viscosity. Rodlike anisotropic mullite grains started to grow in titania-doped diphasic mullite gels once a dense, equiaxed microstructure was achieved. The onset temperature for anisotropic grain growth decreased with increasing titania concentration because the sintering temperature for final-stage densification decreased. The lowest onset temperature for anisotropic grain growth was ∼1500°C in 5 wt% titania-doped mullite. The aspect ratio and area fraction of anisotropic mullite grains increased with higher titania concentration and were strongly dependent on the initial titania particle size. Kinetic studies demonstrated that anisotropic grain growth in titania-doped diphasic mullite gels followed the empirical equation Gn - G 0 n = Kt , with growth exponents of 3 and 6 for the length [001] and thickness [110] directions, respectively. The activation energies for grain growth were 690 kJ/mol for the length and 790 kJ/mol for the thickness directions.  相似文献   

6.
Master sintering curves were constructed for dry-pressed compacts composed of either a nanocrystalline or a microcrystalline ZnO powder using constant heating rate dilatometry data and an experimentally determined apparent activation energy for densification of 268±25 and 296±21 kJ/mol, respectively. The calculated activation energies for densification are consistent with one another, and with values reported in the literature for ZnO densification by grain boundary diffusion. Grain boundary diffusion appears to be the single dominant mechanism controlling intermediate-stage densification in both the nanocrystalline and the microcrystalline ZnO during sintering from 65% to 90% of the theoretical density (TD). Based on both the consistency of the calculated activation energy as a function of density and the narrow dispersion of the sintering data about the master sintering curve (MSC) for the nanocrystalline ZnO, there is no evidence of either significantly enhanced surface diffusion or grain growth during sintering relative to the microcrystalline ZnO. The MSC constructed for the nanocrystalline ZnO was used to design time–temperature profiles to successfully achieve four different target sintered densities on the MSC, demonstrating the applicability of the MSC theory to nanocrystalline ceramic sintering. The most significant difference in sintering behavior between the two ZnO powders is the enhanced densification in the nanocrystalline ZnO powder at shorter times and lower temperatures. This difference is attributed to a scaling (i.e., particle size) effect.  相似文献   

7.
A calcium magnesium aluminosilicate-based glass that contained 10 wt% of silicon carbide whiskers (SiC w ) as reinforcement was prepared by tape casting, followed by sintering either in a conventional furnace or in a microwave oven. The results were consistent with retardation of glass sintering through whisker bridging. The glass, by itself, was sintered to almost-full density at 750°C for 4 h by conventional furnace sintering; the best sintered composite, with an estimated density of ∼90%, was obtained at 800°C with a dwell time of 4 h. Sintering at a temperature of >800°C did not improve the densification but rather resulted in severe whisker oxidation. A reduced densification rate was observed for the samples that were sintered in nitrogen. By contrast, in the microwave oven, almost-full density for the glass and ∼95% of the theoretical density for the composite were obtainable at 850°C for 15 min, which represented a reduction of ∼10 h of the total processing time and a reduced SiC w oxidation.  相似文献   

8.
Nanocrystalline, single-phase undoped ZnO was sintered to 95%–98% of theoretical density at 650°–700°C, using pressureless isothermal sintering. The density increased very rapidly at 500°–600°C, remained constant with sintering temperature until ∼900°C, and then decreased slightly. The estimated activation energy for densification at 600°–700°C (275 kJ/mol) was comparable to grain-growth activation energies previously reported for microcrystalline ZnO but much greater than the grain-growth activation energy measured in the present work. A bimodal microstructure, consisting of nanocrystalline grains within larger ensembles ("supergrains"), was observed, and both modes grew as the sintering temperature increased. The grain-growth activation energy for the nanocrystalline grains was extremely low, ∼20 kJ/mol. The activation energy for the growth of the supergrains depended strongly on temperature but was ∼54 kJ/mol at >500°C. The important mechanisms probably are rearrangement of the nanoparticle grains, with simultaneous surface and boundary diffusion, and vapor transport above 900°C.  相似文献   

9.
Borosilicate glass powder with simulated high-level radioactive waste (HLW) was solidified by a hydrothermal hotpressing method. A leach test of the solid bodies was performed. Twenty-one elements composing the simulated HLW in the leachant were analyzed quantitatively by neutron activation analysis. The addition of Al(OH)3, up to 7.5% of the starting powder, reduced the leachability. The first densification of the starting powder during hydrothermal hot-pressing takes place by the rearrangement of particles, followed by solution-precipitation. The densification fits two types of rate equations. Activation energies obtained from two rate equations agree and are smaller than those of the glass without simulated HLW. The results indicate that simulated HLW in the starting powder acts to reduce the activation energy for densification during hydrothermal hot-pressing.  相似文献   

10.
Densification of alumina coated with a MgO-Al2O3-SiO2 (MAS) glass was investigated from 1400° to 1460°C. Spinel was observed to form at the liquid-alumina interface, whereas mullite crystallized uniformly in the liquid. Spinel and mullite crystallization kinetics were accelerated by smaller alumina particle size. Mullite and spinel crystallization retarded densification by forming a percolating network. Boron doping suppressed spinel formation, and thus alumina sintered to higher densities at low temperature. The concept of glass basicity is proposed as a useful guide for selecting dopants for low-temperature sintering.  相似文献   

11.
Si3N4 compacts, containing ≅7 wt% of both BeSiN2 and SiO2 as densification aids, can be reproducibly sintered to relative densities >99% by a gas-pressure sintering process. Nearly all densification takes place via liquid-phase sintering of transformed β-Si3N4 grains at T =1800° to 2000°C. Compacts with high density are produced by first sintering to the closed-pore stage (≅92% relative density) in 2.1 MPa (20 atm) of N2 pressure at 2000°C and then increasing the N2 pressure to 7.1 MPa (70 atm) where rapid densification proceeds at T = 1800° to 2000°C. The experimental density results are interpreted in terms of theoretical arguments concerning the growth (coalescence) of gas-filled pores and gas solubility effects. Complex chemical reactions apparently occur at high temperatures and are probably responsible for incomplete understanding of some of the experimental data.  相似文献   

12.
The sintering behavior of powder compacts formed from spherical, nearly monosized SiO2 particles was investigated. Highly ordered compacts sintered to high density and translucency at 1000°C. In contrast, less homogeneous samples prepared from flocculated suspensions remained highly porous after sintering under the same conditions. Densification kinetics were determined over the temperature range 900° to 1050°C for ordered compacts. The viscosity at each sintering temperature and the activation energy for viscous flow were determined using available sintering models. Sintering of ordered compacts is divided into several stages. Densification, mercury porosimetry, and electron microscopy results indicate that the first stage is dominated by the shrinkage of three-particle pore channels, whereas the second stage primarily involves the shrinkage of four-particle pore channels.  相似文献   

13.
The microstructure evolution and densification of alumina containing 10 vol% calcium aluminosilicate glass and 0.5 wt% magnesium oxide sintered at 1600°C were quantified by measuring the evolution of pore-size distribution, the redistribution of liquid phase, and the fraction of closed and open pores. The densification stopped at a limiting relative density during the final stage of sintering, and the small and large pores were filled simultaneously by glass during sintering. In addition, the results indicate that the pressure build-up of the trapped gases in pores causes a significantly negative contribution to the driving force, and consequently the observed reduction in densification during the final stage of liquid phase sintering.  相似文献   

14.
Densification during liquid-phase sintering of WC–Co with various WC powder sizes has been measured in order to identify the densification mechanism. During heating of powder compacts in the solid state, densification was enhanced with a reduction of WC powder size. However, the behavior was reversed when the densification occurred in the presence of a liquid: enhanced densification with increasing WC powder size. This result is in contradiction to a prediction of the conventional theory of liquid-phase sintering, the contact flattening theory, but in good agreement with a prediction of the pore-filling theory. Microstructural analysis further confirmed that the densification at the liquid-phase sintering temperature occurred by pore filling. The calculated densification kinetics based on the pore-filling theory also fitted well with the measured data. The observed densification behavior thus demonstrates experimentally the prediction of the pore-filling theory that the densification is enhanced with increasing average grain size for the same pore size distribution.  相似文献   

15.
The fabrication temperature was the principal variable in a kinetic study of the densification of hafnium diboride in high-pressure hot-pressing. Densification studies for conventional hot-pressing were reviewed and correlated with the high-pressure hot-pressing results. The consolidation of HfB2 in the open pore region during high-pressure hot-pressing is attributed to particle rearrangement caused by grain boundary sliding and fragmentation. The final stage of densification (relative density >90%) was analyzed in terms of the Nabarro-Herring vacancy creep model. An activation energy of 22,900 cal/mole was obtained for the rate-controlling step in the creep process.  相似文献   

16.
The kinetics of the final stage of densification of fine-grained aluminum oxide were studied by vacuum hot-pressing between 1150° and 1350°C and from 2000 to 6000 psi. The kinetics are consistent with the Nabarro-Herring diffusional creep model. The activation energy for the final stage densification is 115 kcal/mole which agrees with the activation energy for diffusion of aluminum ions. Densification takes place by particle rearrangement and by diffusional creep. The extent of densification in the initial stage depends on the amount of sliding, fragmentation, and plastic flow. The final stage of densification takes place by diffusional creep and is controlled by aluminum ion diffusion in aluminum oxide. It is shown that gases entrapped within pores of the hot-pressed compact will produce end-point porosities. A technique is described for removing adsorbed water when it is the source of the Presented at the Fall Meeting of the Basic Science Division and the Seventeenth Pacific Coast Regional Meetings of the American Ceramic Society, San Francisco, California, October 30, 1964.  相似文献   

17.
Microwave Sintering of Alumina at 2.45 GHz   总被引:3,自引:0,他引:3  
The sintering kinetics and microstructural evolution of alumina tubes (∼17 mm length, ∼9 mm inner diameter, and ∼11 mm outer diameter) were studied by conventional and microwave heating at 2.45 GHz. Temperature during microwave heating was measured with an infrared pyrometer and was calibrated to ±10°C. With no hold at sintering temperature, microwave-sintered samples reached 95% density at 1350°C versus 1600°C for conventionally heated samples. The activation energy for microwave sintering was 85 ± 10 kJ/mol, whereas the activation energy for conventionally sintered samples was 520 ± 14 kJ/mol. Despite the difference in temperature, grains grew from ∼1.0 μm at 86% density to ∼2.6 μm at 98% density for both conventionally sintered and microwave-sintered samples. The grain size/density trajectory was independent of the heating source. It is concluded that the enhanced densification with microwave heating is not a consequence of fast-firing and therefore is not a result in the change in the relative rates of surface and grain boundary diffusion in the presence of microwave energy.  相似文献   

18.
Grain growth and densification were studied in a hot-pressed ferroelectric composition of 65/35 (Zr/Ti molar ratio) lead zirconate-lead titanate containing 2 at.% bismuth. The grain growth and densification rate processes were measurable from 1050° to 1300°C and from 700° to 1100°C, respectively. Grain growth as a function of time followed a ⅓ power law ( D (grain size) = kt ). An activation energy of 95 kcal/mole was calculated. Densification was a two-stage process as a function of time: a rapid initial stage and a slower final stage. The initial stage of densification behaved viscously with an activation energy of 36.7 kcal/mole. Grain boundary sliding and the Nabarro-Herring mechanism of stress-directed movement of vacancies are suggested as the densification mechanisms. Results indicated that densification is sensitive to stoichiometry and to additives.  相似文献   

19.
Effects of a liquid-phase-sintering aid, BaCuO2+ CuO (BCC), on densification and microwave dielectric properties of (Zr0.8Sn0.2)TiO4 (ZST) ceramics have been investigated. The densification kinetics of ZST are greatly enhanced with the presence of 2.5–5 wt% BCC, but become retarded when the amount of BCC increases further. At a given BCC content, moreover, slower densification kinetics are observed with a larger particle size of ZST. The above results are attributed to a chemical reaction taking place at the interface of BCC/ZST during firing. The ZST dissolves into BCC, forming crystalline phases of ZrO2, SnO2, CuO, and BaTi8O16 which reduce the amount of BCC flux available for liquid-phase sintering. The crystallization kinetics become more significant, compared with densification kinetics, with increasing the amount of BCC and the particle sizes of ZST. For samples with 2.5–5 wt% BCC, a high relative sintered density is obtained at 1000°C and the resulting microwave ceramics have a dielectric constant and a value of Q at 7 GHz in the ranges of 35–38 and 2800–5000, respectively.  相似文献   

20.
Effect of Green Density on Densification and Creep During Sintering   总被引:4,自引:0,他引:4  
The effect of green density on both the densification rate and the creep rate was measured simultaneously during sintering by loading dilatometry. The experiments were performed on zinc oxide powder compacts with five different green densities covering a range of 0.39 to 0.73 of theoretical. The samples were heated at a constant rate of 4°C/min up to 1100°C in air. The densification rate at any temperature increases significantly with decreasing green density. The data for the densification rate and creep rate as a function of density show two quite distinct regimes of behavior; the rates were strongly dependent on density below 0.80, while above this value they were weakly dependent on density. The ratio of the densification rate to the creep rate was almost independent of temperature but increased almost linearly with increasing green density. The representation of the data in terms of models for sintering and creep is discussed.  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