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 234 毫秒
1.
The surface of an aluminosilicate ceramic was transformed to a corundum layer of high specific surface area by heating at 1300°–1450°C in a controlled reducing atmosphere. This procedure selectively reduced and volatilized the silica of the glass and mullite, and the alumina of the mullite formed a layer of corundum crystals with a thickness of ≤20 μm and a specific surface area of ≤16 m2/g. Specific surface area remained stable at 10.5 m2/g after prolonged heating at 1300°C in air, and at 8.5 m2/g at 1450°C. These materials are well suited for use as catalyst supports in applications such as catalytic combustion at temperatures in this range.  相似文献   

2.
The effects of tartaric acid that was used in sol–gel processing on specific surface area, pore-size distribution, and gel structure of sol–gel-derived silica gels were investigated. The specific surface area of silicas that were calcined at 450°C increased from ∼600 m2/g to ∼1200 m2/g as the amount of tartaric acid that was used increased. The pore-size distribution changed as the surface area increased, and only the gels that had a surface area of 930-990 m2/g showed a very sharp pore-size distribution in the mesopore range. The difference in gel structure and properties was explained in terms of the acidity of tartaric acid and the inhibition of condensation among primary particles through the coordination or adsorption of tartaric acid on the particle surface. From the thermal behavior of the gels with different features, it was concluded that gel properties are determined not only by the structure of the precursor gel but also by the surface activity. The gel with uniform mesopores gave a high surface area over a wide range of calcination temperatures.  相似文献   

3.
Aluminum nitride (AlN)–silicon carbide (SiC) nanocomposite powders were prepared by the nitridation of aluminum-silicon carbide (Al4SiC4) with the specific surface area of 15.5 m2·g−1. The powders nitrided at and above 1400°C for 3 h contained the 2H-phases which consisted of AlN-rich and SiC-rich phases. The formation of homogeneous solid solution proceeded with increasing nitridation temperature from 1400° up to 1500°C. The specific surface area of the AlN–SiC powder nitrided at 1500°C for 3 h was 19.5 m2·g−1, whereas the primary particle size (assuming spherical particles) was estimated to be ∼100 nm.  相似文献   

4.
Crystalline lithium metasilicate (Li2SiO3) nanoparticles have been synthesized using a sol–gel process with tetraethylorthosilicate and lithium ethoxide as precursors. The particle size examined by using transmission electron microscopy and BET-specific surface area techniques is in the range 5–50 nm, depending on the temperature at which the material is calcined. The crystalline Li2SiO3 forms at ambient temperature (∼40°C), and it remains in this phase after calcination at temperatures up to 850°C. The BET-specific surface area is ∼110 m2/g for material calcined at temperatures below 500°C, decreasing to ∼29 and ∼0.7 m2/g following calcination at 700° and 850°C, respectively. Solid-state 29Si NMR spectroscopy shows the presence of only Q 2 structural units in the material. The lithium metasilicate is further characterized using differential scanning calorimetry and thermogravimetric analysis, and Fourier transform infrared spectroscopy.  相似文献   

5.
Phase-pure brookite of high crystallinity, which was classically obtained via hydrothermal treatment, has been synthesized under ambient pressure at 70°C via reacting a mixed solution of urea and titanium (III) chloride (instead of the widely used titanium (IV) compounds). The resultant particles are monodispersed spheres (∼154 nm) composed of brookite nanocrystals (∼25 nm), which are stable in terms of phase purity and morphology up to ∼500°C, above which a direct transition to rutile occurred. The as-made powder has a high specific surface area of ∼41.2 m2/g, which rapidly decreased to ∼9.7 m2/g after transforming to rutile at 700°C. The brookite powder shows good catalytic property for the decomposition of acetaldehyde under UV radiation.  相似文献   

6.
Differential scanning calorimetry has been used to make direct measurements of the excess enthalpy of TiO2 (rutile) with an initial grain size of 30–70 nm. When the heat released during grain growth is normalized to the change in grain boundary area, the specific excess enthalpy at low temperatures and fine grain sizes (600°–780°C, 30–200 nm) is found to be 0.5–1 J/m2, while values averaged over a larger temperature and size range (600°–1300°C, 30 nm- ∼2 μm) are 1.3–1.7 J/m2. After exclusion of extraneous contributions from other heat-dissipating processes, origins of a specific grain boundary enthalpy that increases with grain size or temperature are considered, including solute segregation, changes in grain boundary structure, and contributions from grain boundary triple junctions. It is concluded that the most plausible explanation is a size-dependent nonstoichiometry of TiO2 due to the impingement of space charge layers in the temperature and grain size range of the experiments.  相似文献   

7.
Porous Silicon Oxycarbide Glasses   总被引:1,自引:0,他引:1  
High-surface-area silicon oxycarbide gels and glasses were synthesized from mixtures of methyldimethoxysilane (MDMS) and tetraethoxysilane (TEOS) through acidic hydrolysis and condensation. A surface area of ∼275 m2/g and an average pore size of ∼30 Å was obtained for a 50% MDMS-50% TEOS glass at 800°C under a flowing argon atmosphere. The average pore size was increased by aging the precursor gels in ammonium hydroxide. The increased average pore size and the higher strength of the mesoporous gel network enhanced the surface-area stability of the glasses; in this case, surface areas >200 m2/g were retained at 1200°C under an argon atmosphere. 29Si MAS NMR spectra revealed that an oxycarbide structure was established in the mesoporous glasses obtained after pyrolysis of the aged gels. The role of carbon was demonstrated by comparing the surface-area stability of the oxycarbide glasses with that of pure silica and that of oxycarbide glasses where all the carbon groups were removed through low-temperature plasma-oxidation treatments. In the absence of carbon, the thermal stability of the surface area decreased dramatically.  相似文献   

8.
The influence of magnesium, phosphorus, and iron additions on the low-temperature (≤1000°C) sintering of nanocrystalline α-Al2O3 derived from α-AlOOH has been investigated. α-AlOOH powder with a surface area of 50 m2/g yielded α-Al2O3 products with surface areas of 150 and 80 m2/g after dehydration at temperatures of 400° and 500°C, respectively. However, these products were prone to sintering at >600°C, and the surface area was reduced to 15 m2/g within only 1 h at 1000°C. Although magnesium and iron doping had no discernible effect, the presence of phosphorus inhibited sintering and surface-area loss significantly. Samples doped with 1%–2% phosphorus had surface areas of >31 m2/g after 100 h at 1000°C. Atomic force microscopy studies of α-Al2O3 pseudomorphs derived from α-AlOOH single crystals also demonstrated the inhibiting effect of phosphorus, as the rate of crack elimination was reduced on phosphorus-modified surfaces. The effects of the dopants are discussed with regard to their potential influence on α-Al2O3 surface energy and diffusivity.  相似文献   

9.
Zirconium diboride (ZrB2) was densified (>98% relative density) at temperatures as low as 1850°C by pressureless sintering. Sintering was activated by removing oxide impurities (B2O3 and ZrO2) from particle surfaces. Boron oxide had a high vapor pressure and was removed during heating under a mild vacuum (∼150 mTorr). Zirconia was more persistent and had to be removed by chemical reaction. Both WC and B4C were evaluated as additives to facilitate the removal of ZrO2. Reactions were proposed based on thermodynamic analysis and then confirmed by X-ray diffraction analysis of reacted powder mixtures. After the preliminary powder studies, densification was studied using either as-received ZrB2 (surface area ∼1 m2/g) or attrition-milled ZrB2 (surface area ∼7.5 m2/g) with WC and/or B4C as a sintering aid. ZrB2 containing only WC could be sintered to ∼95% relative density in 4 h at 2050°C under vacuum. In contrast, the addition of B4C allowed for sintering to >98% relative density in 1 h at 1850°C under vacuum.  相似文献   

10.
Lanthanum zirconate (La2Zr2O7) was prepared by coprecipitating lanthanum nitrate and zirconyl oxychloride at pH 10, followed by ethanol washing. The initial high surface area of ∼304 m2·g−1 decreased very rapidly with increased sintering temperature and decreased to an immeasurably small value after heating at 1200°C for 15 h. The major parameters studied were phase evolution, crystallite size, porosity, surface area reduction, and shrinkage during sintering. Three temperature regions were identified based on these studies: below the crystallization temperature, between the crystallization temperature and ∼1100°C, and above 1100°C. The main contribution of surface area reduction in the region 800°–1100°C was due to surface diffusion; the main contribution above 1100°C was due to grain-boundary diffusion coupled with surface diffusion.  相似文献   

11.
The effect of attrition milling on the solid-state oxidation of aluminum powder is important for the reaction-bonded aluminum oxide process. Attrition milling increased the surface area to 14.4 and 20.2 m2/g versus 1.2 m2/g for unmilled powder and smeared the Al particles, and the surface was hydrolyzed to form bayerite and boehmite. Upon heating the hydroxides decompose to form an 11–13 nm thick amorphous plus γ-Al2O3 layer which subsequently retards oxidation kinetics. The oxidation per unit area decreases for the higher surface area powders at temperatures below the critical temperature but the total oxidation of the milled powder is ∼70% versus ∼9% for the as-received powder because of the higher surface area. The critical temperature depends on Al particle surface characteristics and is defined as the transition temperature above which the initial rate of oxidation is linear, not parabolic. Above the critical temperature the oxidation per unit area decreases significantly. In addition, linear oxidation occurs faster than parabolic oxidation and thus the initial fast oxidation kinetics (i.e., linear) can cause thermal runaway during oxidation. Therefore, oxidation below the critical temperature is essential to maximize solid-state oxidation and to prevent thermal runaway. The critical temperatures for the as-received (1.24 m2/g), the 6 h (14.4 m2/g), and 8 h (20.2 m2/g) attrition-milled Al powders were 500°, 475°, and 500°C, respectively. A model for oxidation during the parabolic and linear oxidation stages is presented.  相似文献   

12.
Addition of α-Fe2O3 seed particles to alkoxide-derived boehmite sols resulted in a 10-fold increase in isothermal rate constants for the transformation of γ- to α-Al2O3. Changes in porosity and surface area with sintering temperature showed no effect of seeding on coarsening of the transition alumina gels, but the 200-fold decrease in surface area associated with transformation to α-Al2O3 occurred ∼ 100°C lower in seeded gels compared with unseeded materials. As a result of high nucleation frequency and reduced microstructure coarsening, fully transformed seeded alumina retained specific surface areas >22 m2/g and exhibited narrow pore size distributions, permitting development of fully dense, submicrometer α-Al2O3 at ∼ 1200°C.  相似文献   

13.
Dense Nb2AlC ceramic was synthesized from NbC, Nb, and Al powder mixture at 1650°C and a pressure of 30 MPa for 90 min using an in situ reaction/hot-pressing method. The reaction kinetics, microstructure, physical, and mechanical properties of the fabricated material were investigated. A thermal expansion coefficient of ∼8.1 × 10−6 K−1 was measured in the temperature range of 30°–1050°C. At room temperature a thermal conductivity of ∼20 W·(m·K)−1 and a Vickers hardness of ∼4.5 GPa were determined. The material attained Young's modulus, four-point bending strength and fracture toughness of ∼294 GPa, ∼443 MPa, and ∼5.9 MPa·m1/2, respectively. The nanolayered grains with a mean grain size of 17 μm contributed to the damage tolerance of this ceramic. Quenching from 600°, 800°, and 1000°C into water at room temperature resulted in decrease in bending strength from 443 MPa for the as-synthesized Nb2AlC to 391, 156, and 149 MPa, respectively.  相似文献   

14.
Aqueous Combustion Synthesis of Strontium-Doped Lanthanum Chromite Ceramics   总被引:2,自引:0,他引:2  
An aqueous combustion synthesis is used to produce powders of La0.8Sr0.2CrO3 perovskite. It is shown that interaction between chromium nitrate and glycine controls the process. In addition, it is suggested that glycine reacts with products of nitrate decomposition to yield an intermediate compound, which decomposes exothermically providing high-temperature conditions for complex oxide formation. It is remarkable that although reaction temperature is high (up to 800°C) and characteristic time is small (∼1 s) for synthesis under the self-propagating high-temperature mode, the produced perovskites have high specific surface area (∼40 m2/g) and well-defined crystalline structure. As a result, ceramics sintered by using these powders are dense (∼96% of theoretical) and possess high electronic and low ionic conductivities, important for interconnect applications in solid oxide fuel cells.  相似文献   

15.
Porous Cr3C2 grains (∼300 to 500 μm) with ∼10 wt% of Cr2O3 were prepared by heating a mixture of MgCr2O4 grains and graphite powder at 1450° to 1650°C for 2 h in an Al2O3 crucible covered by an Al2O3 lid with a hole in the center. The porous Cr3C2 grains exhibited a three-dimensional network skeleton structure. The mean open pore diameter and the specific surface area of the porous grains formed at 1600°C for 2 h were ∼3.5 (μm and ∼6.7 m2/g, respectively. The present work investigated the morphology and the formation conditions of the porous Cr3C2 grains, and this paper will discuss the formation mechanism of those grains in terms of chemical thermodynamics.  相似文献   

16.
Silica-modified titanium oxide (S-TiO2) powders that have an anatase structure were synthesized via the thermal decomposition of mixtures of titanium(IV) isopropoxide (TIP) and tetramethyl orthosilicate (TMOS) in toluene at 300°C. These S-TiO2 materials had high rutile-transformation temperatures and maintained large surface areas at elevated temperatures (550°–1000°C). For example, the product that was prepared from a 9:1 TIP:TMOS mixture transformed to rutile at ∼1100°C and possessed a surface area of 160 m2/g, even after calcination at 800°C for 1 h.  相似文献   

17.
Compression creep tests were performed on fully dense specimens of UC1.01, UC1.05, UC1.01.+ 4 wt% W, and U0.9Zr0.1C1.01+ 4 wt% W. Steady-state creep rates were measured from 1400° to 1800°C in a vacuum of 1.33 × 10-3 N/m2 (1 × 10-5 torr) at stresses of 4.55 to 69.0 MN/m2 (660 to 10,000 psi). The data for UC1.01 could best be fit by an expression of the form ɛ= 1773σ6.024 exp (106.5/RT) , where σ is the steady-state creep rate (h-l), σ is the applied stress (MN/m2), and the creep activation energy is given in kcal/mol. The stress dependence for creep of UC1.05 decreased with decreasing temperature because of second-phase precipitation; therefore, a unique creep activation energy could not be established for this U/C ratio. At all temperatures, the creep strength of UC1.05 exceeded that of UC1.01. For example, at 1700 ° C steady-state creep rates for UC1.05 are ∼1/4 those for UC1.01, but at 1400°C the creep rates are ∼ 3 orders of magnitude less. At 1700°C, creep rates for UC alloys are ∼4 orders of magnitude lower than those for unalloyed UC1.01.  相似文献   

18.
The changes in surface area and mesoporosity in aggregates of ∼0.01 μm cross-section CaO particles when heated in CO2at 686°C were determined from N2 adsorption isotherms. Initially, the surface area decreases rapidly with little change in porosity. When the surface area has decreased below ∼90 m2/g, surface area and porosity variations become consistent with expectations for coarsening by grain-boundary or bulk diffusion. The initial rapid decrease in surface area must result from CO2-catalyzed surface diffusion, but the data suggest that surface diffusion is not rate-limiting. The rate-limiting step may be reaction of CO2 to form surface CO32- ions or decomposition of these ions to O2- ions and CO2 gas.  相似文献   

19.
Tetragonal Zirconia Polycrystals Reinforced with SiC Whiskers   总被引:2,自引:0,他引:2  
The microstructure and the mechanical properties of hot-pressed tetragonal ZrO2 polycrystals (TZP) reinforced with up to 30 vol% SiC whiskers were studied. The SiC whisker-TZP composites were stable under the hot-pressing conditions at 1450°C. Annealing in an oxidizing atmosphere at ∼1000°C resulted in glass formation and microcracking caused by whisker oxidation and transformation of the ZrO2 grains near the whiskers to monoclinic symmetry. The fracture toughness was markedly improved by the dispersed whiskers (∼12 Mpa·m1/2 at 30 vol% SiC) compared to the values measured for the matrix (∼6 Mpa·m1/2). The flexural strength of the hot-pressed TZP-30 vol% SiC whisker composite at 1000°C (∼400 MPa) was twice that of the TZP matrix.  相似文献   

20.
Ultrafine ZrB2–SiC composite powders have been synthesized in situ using carbothermal reduction reactions via the sol–gel method at 1500°C for 1 h. The powders synthesized had a relatively smaller average crystallite size (<200 nm), a larger specific surface area (∼20 m2/g), and a lower oxygen content (∼1.0 wt %). Composites of ZrB2+20 wt% SiC were pressureless sintered to ∼96.6% theoretical density at 2250°C for 2 h under an argon atmosphere using B4C and Mo as sintering aids. Vickers hardness and flexural strength of the sintered ceramic composites were 13.9±0.3 GPa and 294±14 MPa, respectively. The microstructure of the composites revealed that elongated SiC grain dispersed uniformly in the ZrB2 matrix. Oxidation from 1100° to 1600°C for 30 min showed no decrease in strength below 1400°C but considerable decrease in strength with a rapid weight increment was observed above 1500°C. The formation of a protective borosilicate glassy coating appeared at 1400°C and was gradually destroyed in the form of bubble at higher temperatures.  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