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 critical zeta potential characterises the flocculated-dispersed state transition of a colloidal dispersion. For many colloidal dispersions, yield stress displayed a linear relationship with the square of zeta potential, indicating that they obeyed the Derjaguin-Landau-Verwey-Overbeek (DLVO) theory. From this relationship, the critical zeta potential is obtained from the intercept at the zeta potential axis at a yield stress of zero. The critical zeta potential is a measure of the repulsive potential required to exactly counter the maximum attractive potential between particles in dispersion in the flocculated state. When the forces of interaction between particles in the dispersion are only the van der Waals and electrostatic forces, then the critical zeta potential is indirectly a measure of the van der Waals attractive potential and, hence, it may be used to determine the Hamaker constant of solids in water. This potential is proportional to the square root of the solids Hamaker constant in water. At present, only the ratio of Hamaker constant between two oxides was obtained and compared with that obtained by other techniques. These oxides were ultrapure anatase TiO2 and γ-Al2O3, and they displayed a linear relationship between yield stress and the square of zeta potential. At the conductivity (or ionic strength) of about 3000 μS/cm, the critical zeta potential for both TiO2 and Al2O3 is ∼47 and ∼32 mV, respectively. These critical zeta potential data give a value of 2.2 for the ratio of Hamaker constant of anatase TiO2/H2O/TiO2 to γ-Al2O3/H2O/γ-Al2O3. This ratio compares well with a value ranging from 1.0 to 2.18 for rutile TiO2/H2O/TiO2 to α-Al2O3/H2O/α-Al2O3 where their Hamaker constants were calculated from the Lifshitz theory using full optical spectral data.  相似文献   

2.
The yield stress-DLVO force relationship is obeyed by α-Al2O3 and alumina-coated TiO2 dispersions with adsorbed polyacrylate only if the yield stress and its corresponding zeta potential data were collected in the positively charged region. In this region, the underlying surface positive charge density of the particles exceeds the negative charge density of the polyacrylate. At this state the adsorbed polyelectrolyte lies flat on the particle surface forming a steric layer of fixed thickness at a given polymer concentration. In the negative charge region, the steric layer thickness is not constant and hence yield stress-DLVO relationship is not obeyed. The (critical) zeta potential at the flocculated-dispersed transition state decreases with increasing polymer concentration. This result reflects a decreasing van der Waals force as the steric layer increases in thickness. A steric layer ensured that the surface or zeta potential is sufficiently low in the flocculated regime for the DLVO theory to remain valid. The ratio of the critical zeta potential square between alumina-coated TiO2 and α-Al2O3 is an indication of their Hamaker constants ratio in water. The effect of alumina coating on the value of this ratio is presented and discussed.  相似文献   

3.
Single-crystal α-Al2O3 hexagonal flakes with a diameter of about 200 nm and 20 nm in thickness were obtained by mixing different molar ratios of potassium sulfate to boehmite and heating at 1000 °C. Co-doping 1 mol% TiO2 can increase the shape anisotropy of α-Al2O3 hexagonal flakes, increasing the diameter to 400 nm. The effects of potassium sulfate, Fe2O3 and TiO2 on the phase transformation and morphology development of alumina were investigated using X-ray diffraction analysis (XRD), differential thermal analysis (DTA) and transmission electron microscopy (TEM). The results indicate that co-doping potassium sulfate, Fe3+ and Ti4+ can promote γ → α-Al2O3 phase transformation and change the morphology from a vermicular structure into hexagonal platelets. The shape anisotropy of α-Al2O3 hexagonal flakes can be increased by adding TiO2 due to the segregation of Ti4+ ions onto the surfaces of basal planes of α-Al2O3 single crystal particle.  相似文献   

4.
Crystal-growth-related microstructures and the length-to-diameter ratio of a single-crystal-type α-Al2O3 nanofiber were examined using HR-TEM techniques. The fibers exhibited diameters ranging from 50 to 100 nm and lengths of several tens of micrometers. During thermal treatments, the alumina fiber went through phase transformations similar to boehmite. Therefore, the phase evolution, especially the final θ- to α-Al2O3 stage of the phase transformation, may be the determining factor in the microstructural evolution of the nanofibers. HR-TEM techniques were utilized to demonstrate that the single crystals were formed by the coalescence of well-elongated α-Al2O3 colonies. The fibers grew in the [1 1 0] or [1 1 2] direction instead of [0 0 1]. A thermodynamic analysis revealed that if the α-Al2O3 nanofiber that transformed from θ-Al2O3 behaved in a stable manner, there could be a size ratio limit for the length and diameter of each α-Al2O3 colony. The smallest potential diameter was calculated to be around 17 nm.  相似文献   

5.
We present here new cellular ceramics, fibers and microspheres produced by novel and general microshaping techniques employing colloidal dispersions in solvent free radiation curable monomer mixtures. High loading, low viscosity dispersions of functional nanoparticles such as TiO2 and α-Fe2O3 and colloidal Fe(C2O4)·2H2O metal salt are achieved with comb-polyelectrolyte surfactants and/or mildly polar resins. TiO2 dispersions are spun and solidified “on the fly” by UV radiation into continuous ceramic/polymer nanocomposite fibers with dimensions below 10 μm. Fe(C2O4)·2H2O dispersions provide a UV curable alternative to α-Fe2O3 dispersions which can only be thermally cured. The Fe(C2O4)·2H2O nanocomposites transform to α-Fe2O3 below 550 °C. Novel cellular Al2O3, TiO2 and α-Fe2O3 articles with porosity >80% and precise replication of the pore-forming agents have also been produced from such dispersions. Al2O3 nanocomposite microspheres are produced by emulsification of the dispersions in an appropriate medium and UV curing.  相似文献   

6.
In this work, α-Al2O3 (1–40 vol%) particles (350 nm) were added to a 19.58Li2O·11.10ZrO2·69.32SiO2 (5 µm) glass-ceramic matrix to prepare composites with the purpose of studying the influence of α-Al2O3 on their mechanical, thermal and electrical properties in order to obtain materials for LTCCs applications. The composites, sintered between 800 and 950 °C for 30 min, with relative densities between 85% and 93%, showed zircon and β-spodumene as main crystalline phases. The maximum bending strength (290 MPa) was achieved for composites with 1% α-Al2O3. For composites containing between 1% and 10% α-Al2O3, sintered at 900 °C/30 min, the electrical and thermal conductivities and CTE varied between 3.35 and 1.21×10−10 S/cm, 4.65 and 2.98 W/m K, 9.54 and 3.36×10−6 °C−1, respectively.  相似文献   

7.
Enhanced performance of a macroporous disk alumina support was fabricated through colloidal filtration route, by using α-Al2O3 powder with an average particle size of 1.1 μm. The support, sintered at 1250 °C, showed relative high permeances towards water (101 L h−1 m−2 bar−1) and nitrogen (∼2×10−6 mol m−2 s−1 Pa−1), with an average surface roughness of ∼175 nm and a high mechanical strength of 61.1 MPa. Titania supported γ-Al2O3 mesoporous layers were deposited onto this promising disk α-Al2O3 support through dip-coating. The disk membrane A1100/TiO2/γ-Al2O3, with pore size of ca. 4.4 nm, showed a pure water flux as high as 4.5 L m−2 h−1 bar−1, which is four times higher than that of γ-Al2O3 membrane reported in literature. This mesoporous membrane showed relative high retention rate (∼80%) towards di-valent cations like Ca2+, Mg2+, but not for the mono-valent cation (Na+).  相似文献   

8.
Ni and Pt catalysts supported on α-Al2O3, α-Al2O3-ZrO2 and ZrO2 were studied in the dry reforming of methane to produce synthesis gas. All catalytic systems presented well activity levels with TOF (s−1) values between 1 and 3, being Ni based catalysts more active than Pt based catalysts. The selectivity measured at 650 °C, expressed by the molar ratio H2/CO reached values near to 1. Concerning stability, Pt/ZrO2, Pt/α-Al2O3-ZrO2 and Ni/α-Al2O3-ZrO2 systems clearly show lower deactivation levels than Ni/ZrO2 and Ni or Pt catalysts supported on α-Al2O3. The lowest deactivation levels observed in Ni and Pt supported on α-Al2O3-ZrO2, compared with Ni and Pt supported on α-Al2O3 can be explained by an inhibition of reactions leading to carbon deposition in systems having ZrO2. These results suggest that ZrO2 promotes the gasification of adsorbed intermediates, which are precursors of carbon formation and responsible for the main deactivation mechanism in dry reforming reaction.  相似文献   

9.
α-Al2O3 platelets were prepared by a molten salt synthesis method when NaAlO2 was used as raw material. The effects of the stirring rate during the gel preparation, heating temperature, type and addition amount of molten salts, addition of plate-like α-Al2O3 seeds, additives such as TiOSO4 and Na3PO4·12H2O on the morphology of α-Al2O3 were studied. High stirring rate during the gel preparation and high heating temperature not only help to restrain the overlapping of α-Al2O3 platelets, but also improve the size distribution. When the heating temperature increases to 1200 °C, most of α-Al2O3 platelets are hexagonal in its morphology, and the size of platelets becomes relatively uniform. When Na2SO4-K2SO4 flux is used instead of NaCl-KCl flux, it is easy to obtain α-Al2O3 platelets with a big size. When the molar ratio of salt to final Al2O3 powders increases to 4:1, most of α-Al2O3 platelets are hexagonal, and the overlapping of powders is inhibited. The addition of a small amount of plate-like seeds has a significant effect on the size of α-Al2O3 platelets. With the increase of seed amount, the diameter of α-Al2O3 platelets tends to decrease. The addition of 5.45 wt.% TiOSO4 results in the formation of hexagonal α-Al2O3 platelets with an average diameter of 5.1 μm and an average thickness of 1.4 μm. Thin α-Al2O3 platelets with a discal shape are obtained owing to the co-addition of 0.51 wt.% Na3PO4·12H2O and 3 wt.% TiOSO4.  相似文献   

10.
Fine particles of anatase were suspended in solutions of ammonium alum with Al2O3/TiO2 molar ratios from 0.1:1 to 7:1. By spray drying the suspensions and calcining the spray-dried powders, Al2O3-TiO2 composite particles were obtained. The results show that after the spray drying, coatings of ammomium alum are formed on the surface of the anatase particles, leading to composite precursor powders (CCPs) with larger particle sizes. Upon calcining the CCPs, ammomium alum pyrolyzes to amorphous Al2O3 and anatase transforms into rutile. Both are mainly responsible for the observed particle size reductions as well as the densification of each composite particle. The in-situ formed α-Al2O3 and rutile may have higher reactivities, forming aluminum titanate at 1150 °C, about 130 °C lower than the theoretical temperature for the formation of Al2TiO5 by solid reaction. The reaction between α-Al2O3 and rutile starts from the interface between the anatase and the alum coating and mainly takes place in the single particles formed by spray drying. The molar ratio of Al2O3 to TiO2 influences the final crystalline phases in the composite powders, but not stoichiometrically.  相似文献   

11.
The effects of pH value on the composition, structure, morphology, and phase transformation of aluminum hydroxides prepared by chemical precipitation were studied. Aluminum hydroxide precipitated at the pH values of 5 and 6 is amorphous and transforms to α-Al2O3 at 950 °C via the amorphous aluminum hydroxide → amorphous Al2O3 → α-Al2O3 transformation path. Aluminum hydroxide precipitated at pH = 7 is boehmite and transforms to α-Al2O3 at 950 °C via the γ-AlOOH → γ-Al2O3 → α-Al2O3 path. Aluminum hydroxide precipitated at pH values in the 8 to 11 range is bayerite and transforms to α-Al2O3 at 1000 °C via the α-Al(OH)3 → γ-Al2O3 → ε-Al2O3 + θ-Al2O3 → α-Al2O3 path. Moreover, the pH value affects not only the morphology of aluminum hydroxide particles which changes from ultrafine floccules through 50 nm blowballs then to 150 nm irregular agglomerates with increasing pH value but also the microstructures of final decomposition products of aluminum hydroxides.  相似文献   

12.
Plate-like α-Al2O3 single-crystal particles were successfully synthesized in NaCl–KCl flux using Al(OH)3 powders as starting materials, and the influence of pre-calcining of Al(OH)3 powders on the phase formation and morphology of α-Al2O3 powders was focused. When Al(OH)3 powders are used as starting materials, the synthesized product at 900 °C is mainly composed of α-Al2O3 and κ-Al2O3, and most synthesized particles show alveolate morphology. At 1100 °C, single-phase α-Al2O3 powders are developed, in which there are many aggregations of intensively bound plate-like particles. In contrast, using porous amorphous Al2O3 powders obtained by pre-calcining Al(OH)3 powders at 550 °C for 3 h as the starting material, plate-like α-Al2O3 single-crystal particles can be well developed above 900 °C. The reason of the influence of pre-calcining of Al(OH)3 powders on the phase formation and morphology of α-Al2O3 powders is also discussed in the paper.  相似文献   

13.
FeTi alloy was prepared by a vacuum smelting method, iron titanium oxide nanotube arrays have been made directly by anodization of the FeTi alloy. Morphologies and microstructures of the samples were characterized by scanning electron microscope, transmission electron microscope, and X-ray diffractometer. Influences of temperature and H2O concentration on the morphologies of the nanotube arrays have been discussed in detail. Magnetic properties of the samples have also been investigated. The as-prepared samples were amorphous. When annealed at 500 °C and 550 °C, pesudobrookite Fe2TiO5 was obtained. At 600 °C, there were mixed Fe2TiO5, rutile TiO2, and α-Fe2O3. Magnetic performance of the nanotube arrays exhibited high sensitivity to temperature and changed interestingly upon annealing. The values of the coercivity and remanence were 340 Oe and 0.061 emu/g respectively for the sample annealed at 550 °C.  相似文献   

14.
Ultrafine α-Al2O3 powders were prepared by a gel combustion method and the agglomeration characteristic of the resultant powders was studied. A variety of fine crystallite α-Al2O3 powders with different agglomeration structures could be obtained by altering the citrate-to-nitrate ratio γ and calcining the precursors at 1050 °C for 2 h. All the powders were of nearly equivalent crystallite size (60–80 nm) except for the P1 powder (113 nm) from the gel with γ = 0.033. The primary crystallites of the obtained α-Al2O3 powders were formed into large secondary particles with different degree of agglomeration. Except for the powder P1, the mean particle sizes from specific surface area and particle size distribution measurement increase with increasing citrate-to-nitrate ratio in the fuel-lean condition and decrease in the fuel-rich condition. Densities of alumina ceramics from powders P4 and P5 sintered at different temperatures were relatively low due to the wide particle size distribution.  相似文献   

15.
Al2O3 nanopowders were synthesized by a simple chitosan-polymer complex solution route. The precursors were calcined at 800–1200 °C for 2 h in air. The prepared samples were characterized by XRD, FTIR and TEM. The results showed that for the precursors prepared with pH 3–9 γ-Al2O3 and δ-Al2O3 are the two main phases formed after calcination at 800–1000 °C. Interestingly, when the precursor prepared with pH 2 was used, α-Al2O3 was formed after calcination at 1000 °C, and pure α-Al2O3 was obtained after calcination at 1200 °C. The crystallite sizes of the prepared powders were found to be in the range of 4–49 nm, as evaluated by the XRD line broadening method. TEM investigation revealed that the Al2O3 nanopowders consisted of rod-like shaped particles and nanospheres with particle sizes in the range of 10–300 nm. The corresponding selected-area electron diffraction (SAED) analysis confirmed the formation of γ- and α-Al2O3 phases in the samples.  相似文献   

16.
Surface modification and characterization of TiO2 nanoparticles as an additive in a polyacrylic clear coating were investigated. For the improvement of nanoparticles dispersion and the decreasing of photocatalytic activity, the surface of nanoparticles was modified with binary SiO2/Al2O3. The surface treatment of TiO2 nanoparticles was characterized with FTIR. Microstructural analysis was done by AFM. The size, particle size distribution and zeta potential of TiO2 nanoparticles in water dispersion was measured by DLS method. For the evaluation of particle size and the stability of nanoparticles in water dispersions with higher solid content the electroacoustic spectroscopy was made. To determine the applicability and evaluate the transmittance of the nano-TiO2 composite coatings UV–VIS spectroscopy in the wavelength range of 200–800 nm was employed. The results showed that surface treatment of TiO2 nanoparticles with SiO2/Al2O3 improves nanoparticles dispersion and UV protection of the clear polyacrylic composite coating.  相似文献   

17.
J.D.A. Bellido 《Fuel》2009,88(9):1673-1034
ZrO2, γ-Al2O3 and ZrO2/γ-Al2O3-supported copper catalysts have been prepared, each with three different copper loads (1, 2 and 5 wt%), by the impregnation method. The catalysts were characterized by nitrogen adsorption (BET), X-ray diffraction (XRD), temperature programmed reduction (TPR) with H2, Raman spectroscopy and electronic paramagnetic resonance (EPR). The reduction of NO by CO was studied in a fixed-bed reactor packed with these catalysts and fed with a mixture of 1% CO and 1% NO in helium. The catalyst with 5 wt% copper supported on the ZrO2/γ-Al2O3 matrix achieved 80% reduction of NO. Approximately the same rate of conversion was obtained on the catalyst with 2 wt% copper on ZrO2. Characterization of these catalysts indicated that the active copper species for the reduction of NO are those in direct contact with the oxygen vacancies found in ZrO2.  相似文献   

18.
Self ordered arrays of titanium manganese mixed oxide nanotubes were prepared by anodization of Ti8Mn alloy (UNS R56080) under ultrasonication in diluted ethylene glycol containing fluoride. The dimensions of the nanotubes (diameter: 20-100 nm and length: 0.5-2.0 μm) could be tuned by changing the synthesis parameters. The as-anodized nanotubes showed a stoichiometry of (Ti,Mn)O2. Upon annealing at 500 °C in oxygen atmosphere, the nanotubes contained a mixture of anatase + rutile phases of TiO2 and Mn2O3. The composition of the oxide nanotubes was influenced by the chemistry of the phases present in the alloy. More manganese content was observed in the oxide formed on the β-phase than in the oxide layer of α-phase. Anodization in the ultrasonic field increased the kinetics of nanotubular oxide formation and resulted in homogeneous ordering of the nanotubular arrays as compared to the anodization by conventional stirring in the fluoride containing ethylene glycol solution. Whereas, anodization in aqueous acidified fluoride solutions resulted in severe attack of the β-phase and did not show presence of nanotubular oxide structure.  相似文献   

19.
Based on local raw materials, a range of LiZnMg aluminosilicate glasses were prepared to investigate the influence of TiO2, Cr2O3, and ZrO2 on the crystallization behaviour and thermal expansion characteristics. Differential thermal analysis showed that the crystallization propensity increases in the order TiO2 > Cr2O3 > ZrO2. Virgilite, β-spodumene ss, gahnite, enstatite and cristobalite were formed in the prepared glass-ceramics. The microstructure of glass-ceramic samples showed growths of rounded and subrounded grains in the base sample, whereas, somewhat rod-like and accumulated growths appeared in samples containing ZrO2. However, a rather homogeneous texture of accumulated growths was developed in glass-ceramics containing TiO2 and Cr2O3. The coefficient of thermal expansion of parent glasses was sensitive to the type of nucleating agent added (Cr2O3 > TiO2 > ZrO2) varying from 24.8 × 10−7 to 65.1 × 10−7 °C−1 being almost unchanged with the heat-treatment. The microhardness values of glass-ceramic samples were in the 763–779 kg/mm2 range.  相似文献   

20.
The linear and mass ablation rates of Ti2AlC ceramics under an oxyacetylene flame at a temperature up to 3000 °C were examined by measuring the dimensions and weight change of the ablated samples. The linear ablation rate was decreased from 0.14 μm s−1 for the first 30 s of the ablation to 0.08 μm s−1 after 180 s. Ti2AlC ceramics gained small amounts of weight upon ablation, which is attributed to the formation of oxidation products on the ablated surface. The ablation surface exhibits a two-layer structure: an oxide outer layer, consisting mainly of α-Al2O3 and TiO2 and some Al2TiO5, and a porous sub-surface layer containing Ti2Al1−xC and TiCxOy. With increasing ablation time, the content of TiO2 and Al2TiO5 in the outer layer increased, and more pores developed in the sub-surface layer. The thermal oxidation of Ti2AlC under the flame and scouring of the viscous oxidation products by high-speed flow of gas torch are the main ablation mechanisms.  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