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 78 毫秒
1.
The isothermal shrinkage behaviors of fine zirconia powders (containing 2.8–2.9 mol% Y2O3) with specific surface areas of about 6 and 16 m2/g were investigated to clarify the effect of specific surface area on the initial sintering stage. The shrinkage of powder compact was measured under constant temperatures in the range of 1000°–1100°C. The increase in specific surface area enhanced the densification rate with increasing temperature. The values of activation energy ( Q ) and frequency-factor term (β0) of diffusion at initial sintering were estimated by applying the sintering-rate equation to the isothermal shrinkage data. The Q of diffusion changes little but the β0 increases with the increase in specific surface area. It is therefore concluded that the increase in the specific surface area of fine zirconia powder enhances the shrinkage rate because of an increase in the β0 at the initial stage of sintering.  相似文献   

2.
The sintering of dried colloidal SiO2 gels, whose preparation and properties are reported in Part I, is described. The effects of various sintering parameters were studied and the conditions for achievement of the best optical quality include the use of: a pretreatment of the SiO2 at ∼925°C, moderate heating rate (∼400°C/h), He+CI2 atmosphere, 1500° to 1600°C sintering temperature, and 1 to 4 h sintering time. Dynamic sintering kinetic studies (heating rate=400°C/h) show that this SiO2 sinters to nearly theoretical density by about 1380°C. However, optical transparency is achieved by removal of minor residual porosity at above 1500°C. Isothermal sintering data fit to a model assuming interconnecting cylinders of SiO2 predict the proper activation energy for the viscosity if initial stages of sintering are considered. Residual porosity in sintered glasses is related to large interstices in the unsintered gel.  相似文献   

3.
The isothermal shrinkage behavior of 2.9 mol% Y2O3-doped ZrO2 powders with 0–1 mass% Al2O3 was investigated to clarify the effect of Al2O3 concentration on the initial sintering stage. The shrinkage of the powder compact was measured at constant temperatures in the range of 950°–1050°C. The Al2O3 addition increased the densification rate with increasing temperature. The values of apparent activation energy ( nQ ) and apparent frequency-factor term (β0 n ), where n is the order depending on the diffusion mechanism, were estimated at the initial sintering stage by applying a sintering-rate equation to the isothermal shrinkage data. The diffusion mechanism changed from grain-boundary diffusion (GBD) to volume diffusion (VD) by Al2O3 addition and both nQ and β0 n increased with increasing Al2O3 concentration. The kinetic analysis of the sintering mechanism suggested that the increase of densification rate by Al2O3 addition largely depends on the increase of β0 n , that is, the increases of n with GBD→VD change and β0 with an increase in Al2O3 content, although the nQ also increases with Al2O3 addition. This enhanced sintering mechanism is reasonably interpreted by the segregated dissolution of Al2O3 at ZrO2 grain boundaries.  相似文献   

4.
Sinterability of undoped, MgO-doped, and TiO2-doped Al2O3 has been examined by applying reported sintering equations. The order of sinterability was MgO-doped ∼ undoped≪ TiO2-doped Al2O3 in the initial and intermediate stages of sintering, but a relative sintered density at 1600°C for 1 h occurred in the order undoped < TiO2-doped < MgO-doped AI2O3. The dispersion of thermal grooving angles increased in the order MgO-doped < undoped < TiO2-doped Al2O3, The change of sinterability by the dopants is explained in terms of mobility of mass transfer estimated from a densification rate in the initial- and intermediate-stage sintering and of dispersed driving forces of densification and grain growth qualitatively evaluated from the width of the dispersion of thermal grooving angles.  相似文献   

5.
Densification of Alumina and Silica in the Presence of a Liquid Phase   总被引:9,自引:0,他引:9  
Alumina and silica powders were sintered at 1250° to 1460°C in the presence of liquid phases containing CaO, A12O3, and SiO2. The data were analyzed using the equation D = K log t +C and the Arrhenius equation. The activation energy for sintering decreases with increasing amounts of liquid phase.  相似文献   

6.
Solubility of Magnesia in Polycrystalline Alumina at High Temperatures   总被引:1,自引:0,他引:1  
High-purity Al2O3 compacts were doped with 0–350 ppm (by weight) of MgO using a liquid immersion technique and equilibrated at temperatures between 1700° and 2000°C under hydrogen. The solubility limits of MgO in Al2O3 at temperatures of 1720° and 1880°C were very low, ∼75 and 175 ppm, respectively. Variation of MgO solubility with temperature could be represented by the equation, ln Mg/Al = 3.80–2.63 × 104/ T . The small MgO solubilities were understood by the high enthalpy (326 kJ/mol) of solution. The results of this study suggested that previous investigations on sintering and grain-growth mechanisms in MgO-doped Al2O3 were probably not done in single-phase Al2O3 solid solutions. However, the conclusions on sintering and grain-growth mechanisms in prior research work in MgO-doped A2O3 may be correct. The effects of SiO2 impurity and grain size on MgO solubility are discussed. Previous grain-growth experiments in MgO-doped Al2O3 are described that demonstrate the clearest evidence for grain-boundary mobility controlled by a solid-solution mechanism.  相似文献   

7.
Silicon nitride ceramics were prepared by spark plasma sintering (SPS) at temperatures of 1450°–1600°C for 3–12 min, using α-Si3N4 powders as raw materials and MgSiN2 as sintering additives. Almost full density of the sample was achieved after sintering at 1450°C for 6 min, while there was about 80 wt%α-Si3N4 phase left in the sintered material. α-Si3N4 was completely transformed to β-Si3N4 after sintering at 1500°C for 12 min. The thermal conductivity of sintered materials increased with increasing sintering temperature or holding time. Thermal conductivity of 100 W·(m·K)−1 was achieved after sintering at 1600°C for 12 min. The results imply that SPS is an effective and fast method to fabricate β-Si3N4 ceramics with high thermal conductivity when appropriate additives are used.  相似文献   

8.
Small amounts of Li2O result in sintering in the AIN-Y2O3-CaO and AIN-CaO systems at firing temperatures <1600°C. The effect is ascribed to reduction of the liquidus temperature. Furthermore, Li2O is removed by volatization at temperatures from 1300° to 1600°C, and its content decreases several ppm from the initial 0.3 wt%. Li2O-doped AIN specimens containing Y2O3 and CaO additives are well densified by firing at 1600°C for 6 h, and their thermal conductivity is 135 W.m−1.K−1.The effect of Li2O addition on sintering and thermal conductivity also is discussed through thermo-dynamic considerations.  相似文献   

9.
Hafnium diboride (HfB2)- and hafnium carbide (HfC)-based materials containing MoSi2 as sintering aid in the volumetric range 1%–9% were densified by spark plasma sintering at temperatures between 1750° and 1950°C. Fully dense samples were obtained with an initial MoSi2 content of 3 and 9 vol% at 1750°–1800°C. When the doping level was reduced, it was necessary to raise the sintering temperature in order to obtain samples with densities higher than 97%. Undoped powders had to be sintered at 2100°–2200°C. For doped materials, fine microstructures were obtained when the thermal treatment was lower than 1850°C. Silicon carbide formation was observed in both carbide- and boride-based materials. Nanoindentation hardness values were in the range of 25–28 GPa and were independent of the starting composition. The nanoindentation Young's modulus and the fracture toughness of the HfB2-based materials were higher than those of the HfC-based materials. The flexural strength of the HfB2-based material with 9 vol% of MoSi2 was higher at 1500°C than at room temperature.  相似文献   

10.
The results obtained from the sintering of Al2O3–50TiC (in weight percent) composite in the temperature range from 1650° to 1800°C with addition of Y2O3 are presented. Densification is accelerated by the formation of liquid at temperatures above 1750°C, and 99% of theoretical density can be achieved by vacuum sintering at 1800°C for 15 min. The liquid presented at the sintering temperature is crystallized to YAG (Y3Al5O12) during cooling.  相似文献   

11.
Inhibition of cubic-rhombohedral phase transformation and low-temperature sintering at 1000°C were achieved for 10-mol%-Sc2O3-doped cubic-ZrO2 by the presence of 1 mol% Bi2O3. The powders of 1-mol%-Bi2O3–10-mol%-Sc2O3-doped ZrO2 were prepared using a hydrolysis and homogeneous precipitation technique. No trace of rhombohedral-ZrO2 phase could be detected, even after sintering at 1000°–1400°C. The average grain size of the ZrO2 sintered at 1200°C was >2 μm because of grain growth in the presence of Bi3+. Cubic, stabilized Bi-Sc-doped ZrO2 sintered at 1200°C had sufficient conductivity at 1000°C (0.33 S/cm) to be used as an electrolyte for a solid-oxide fuel cell (SOFC) and at 800°C (0.12 S/cm) for an intermediate-temperature SOFC.  相似文献   

12.
The effect of rare-earth oxide additives on the densification of silicon nitride by pressureless sintering at 1600° to 1700°C and by gas pressure sintering under 10 MPa of N2 at 1800° to 2000°C was studied. When a single-component oxide, such as CeO2, Nd2O3, La2O3, Sm2O3, or Y2O3, was used as an additive, the sintering temperature required to reach approximate theoretical density became higher as the melting temperature of the oxide increased. When a mixed oxide additive, such as Y2O3–Ln2O3 (Ln=Ce, Nd, La, Sm), was used, higher densification was achieved below 2000°C because of a lower liquid formation temperature. The sinterability of silicon nitride ceramics with the addition of rare-earth oxides is discussed in relation to the additive compositions.  相似文献   

13.
A type of new low sintering temperature ceramic, Li2TiO3 ceramic, has been found. Although it is difficult for the Li2TiO3 compound to be sintered compactly at temperatures above 1000°C for the volatilization of Li2O, dense Li2TiO3 ceramics were obtained by conventional solid-state reaction method at the sintering temperature of 900°C with the addition of ZnO–B2O3 frit. The sintering behavior and microwave dielectric properties of Li2TiO3 ceramics with less ZnO–B2O3 frit (≤3.0 wt%) doping were investigated. The addition of ZnO–B2O3 frit can lower the sintering temperature of the Li2TiO3 ceramics, but it does not apparently degrade the microwave dielectric properties of the Li2TiO3 ceramics. Typically, the good microwave dielectric properties of ɛr=23.06, Q × f =32 275 GHz, τf = 35.79 ppm/°C were obtained for 2.5 wt% ZnO–B2O3 frit-doped Li2TiO3 ceramics sintered at 900°C for 2 h. The porosity was 0.08%. The Li2TiO3 ceramic system may be a promising candidate for low-temperature cofired ceramics applications.  相似文献   

14.
A low-temperature, single step, reactive sintering method for Pb(Mg1/3Nb2/3)O3 (PMN) and PMN–PbTiO3 (PMN–PT) processing was developed based on the coating of Mg(OH)2 on Nb2O5. This method simplified the processing of PMN and PMN–PT to a single step of heat-treatment and decreased the sintering temperature to 1000°C. It was found that the pyrochlore phase formation reaction at 500°C reduced the particle size to 130 nm. The overlap of the pyrochlor-perovskite phase transformation between 700° and 900°C and the densification process between 800° and 1000°C improved the sintering process. These two factors were the major reasons of the low temperature sintering.  相似文献   

15.
The densification of non-oxide ceramics like titanium boride (TiB2) has always been a major challenge. The use of metallic binders to obtain a high density in liquid phase-sintered borides is investigated and reported. However, a non-metallic sintering additive needs to be used to obtain dense borides for high-temperature applications. This contribution, for the first time, reports the sintering, microstructure, and properties of TiB2 materials densified using a MoSi2 sinter-additive. The densification experiments were carried out using a hot-pressing and pressureless sintering route. The binderless densification of monolithic TiB2 to 98% theoretical density with 2–5 μm grain size was achieved by hot pressing at 1800°C for 1 h in vacuum. The addition of 10–20 wt% MoSi2 enables us to achieve 97%–99%ρth in the composites at 1700°C under similar hot-pressing conditions. The densification mechanism is dominated by liquid-phase sintering in the presence of TiSi2. In the pressureless sintering route, a maximum of 90%ρth is achieved after sintering at 1900°C for 2 h in an (Ar+H2) atmosphere. The hot-pressed TiB2–10 wt% MoSi2 composites exhibit high Vickers hardness (∼26–27 GPa) and modest indentation toughness (∼4–5 MPa·m1/2).  相似文献   

16.
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   

17.
Powder compacts consisting of Al, Al2O3, and ZrO2 were heated by microwave radiation. Tracing the phase evolution during reaction bonding revealed the reaction mechanism. In the case of conventional heating, the compacts expanded slightly at temperatures of <700°C due to Al surface oxidation and expanded sharply at temperatures greater than 700°C as oxidation proceeded from the surface to the interior. Then, the compacts shrank at 1550°C due to sintering. For the case of microwave heating, the compacts expanded at temperatures of <550°C due to the formation of Al3Zr. This Al3Zr formation was caused by the preferential heating of ZrO2 relative to Al and Al2O3 by microwave radiation. Then, Al3Zr was oxidized to form Al2O3 and ZrO2 at temperatures of >1000°C. Finally, the compacts shrank at 1550°C due to sintering, similarly to conventional sintering.  相似文献   

18.
Lead-based piezoelectric ceramics typically require sintering temperatures higher than 1000°C at which significant lead loss can occur. Here, we report a double precursor solution coating (PSC) method for fabricating low-temperature sinterable polycrystalline [Pb(Mg1/3Nb2/3)O3]0.63-[PbTiO3]0.37 (PMN–PT) ceramics. In this method, submicrometer crystalline PMN powder was first obtained by dispersing Mg(OH)2-coated Nb2O5 particles in a lead acetate/ethylene glycol solution (first PSC), followed by calcination at 800°C. The crystalline PMN powder was subsequently suspended in a PT precursor solution containing lead acetate and titanium isopropoxide in ethylene glycol to form the PMN–PT precursor powder (second PSC) that could be sintered at a temperature as low as 900°C. The resultant d 33 for samples sintered at 900°, 1000°, and 1100°C for 2 h were 600, 620, and 700 pm/V, respectively, comparable with the known value. We attributed the low sintering temperature to the reactive sintering nature of the present PMN–PT precursor powder. The reaction between the nanosize PT and the submicrometer-size PMN occurred roughly in the same temperature range as the densification, 850°–900°C, thereby significantly accelerating the sintering process. The present PSC technique is very general and should be readily applicable to other multicomponent systems.  相似文献   

19.
The sintering of ultrafine γ-Al2O3 powder (particle size ∼10–20 nm) prepared by an inert gas condensation technique was investigated in air at a constant heating rate of 10°C/min. Qualitatively, the kinetics followed those of transition aluminas prepared by other methods. Measurable shrinkage commenced at ∼ 1000°C and showed a region of rapid sintering between ∼1125° and 1175°C followed by a transition to a much reduced sintering rate at higher temperatures. Starting from an initial density of ∼0.60 relative to the theoretical value, the powder compact reached a relative density of 0.82 after sintering to 1350°C. Compared to compacts prepared from the as-received powder, dispersion of the powder in water prior to compaction produced a drastic change in the microstructural evolution and a significant reduction in the densification rate during sintering. The incorporation of a step involving the rapid heating of the loose powder to ∼1300°C prior to compaction (which resulted in the transformation to α-Al2O3) provided a method for significantly increasing the density during sintering.  相似文献   

20.
The field-activated sintering technique (FAST) was applied to simultaneously sinter and react sol–gel amorphous powders to form Al2TiO5. Densities close to theoretical and conversion to Al2TiO5 (to 92.5%) have been achieved using FAST at 1050°–1200°C for 10 min. Conventional sintering of the same powders at 1300°C for 2 h resulted in 88.9% Al2TiO5 and ∼75% of theoretical density. The enhanced sintering and compound formation using FAST have been explained by the synergistic effects of precursor reactivity, nanosized powders, and electric-field effects.  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