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 250 毫秒
1.
The effect of autoclaving a zinc oxide preparation containing SO2?4 under 5 and 10 atmospheres is studied by combining X-ray diffraction, differential thermal analysis, thermogravimetry and IR spectroscopy. Textural measurements are also carried out on the parent samples and those produced in the temperature range 200–1000°C.A new phase of a basic carbonates?ulphate, including ammonia in its coordination shell, is observed in the original preparation and having its d distances at 11.060, 8.954 and 2.714 Å. This is transformed to another phase at ~180°C which is also the main phase characterizing the autoclaved samples, and belongs to a basic zinc oxide—sulphate possessing d distances at 7.055, 2.468 and 2.805 Å. Autoclaving the oxide preparation under 10 atm gives hexagonal zinc oxide of high purity and crystallinity at 1000°C. An empirical formula is given for the oxide preparation which describes the different decomposition stages observed. At ~390°C, a reversible reduction process comprising oxygen evolution is observed.Autoclaving increases the area of the parent oxide and at temperatures below 600°C is a function of the structural changes. The autoclaving pressure is insignificant ?600°C.Pore structure analysis showed all the samples to be predominantly mesoporous, coexisting with some micropores except that autoclaved under 5 atm and heated at 250°C which is predominantly microporous. Autoclaving under 5 atm causes narrowing of the pores for products below 600°C. Autoclaving has little effect on the average pore radius ?600°C.Evaluation of the average pore radius from the constructed t-curves for parallel-plate pore idealization is discussed.  相似文献   

2.
The temperatures of phase transitions of the rare earth aluminates have been determined by high temperature X-ray diffractometry. All of the transitions are reversible and occur respectively for Rh ? C at 500°C (LaAlO3), 1330°C (PrAlO3), 1550°C (NdAlO3), and 1950°C (SmAlO3) and for O ? Rh at 770°C (SmAlO3), 1330°C (EuAlO3), and 1700°C (GdAlO3). LnAlO3 perovskites from TbAlO3 up to LuAlO3 are orthorhombic without any phase transition.  相似文献   

3.
Marine fishery wastes such as bivalve shells, crab shells and cuttlebone are rich in calcium. Calcium carbonate derived from these materials can be transformed into calcium oxide by calcination, which is used in a wide variety of applications (e.g., biomaterials for bone and teeth implants and drugs). In this study we analyze the effects of calcination temperatures (550 °C, 700 °C and 900 °C) on characteristics and elemental composition of calcium oxide derived from shells of four marine species collected in Thailand: oyster (Saccostrea cucullata), green mussel (Perna viridis), blue swimming crab (Portunus pelagicus), and cuttlefish (Sepia brevimana). The XRD patterns indicated the complete transformation of calcium carbonate into calcium oxide, observed by the changes of diffraction angles at 900 °C for all calcined samples, except cuttlebone, which was calcined successfully at 700 °C. Likewise, the FT-IR results revealed changes of functional groups at the same calcination temperatures. In addition, ICP-OES showed the effects of calcination temperature on elemental contents: major elements (Ca, P and K) increased in all samples, and some minor elements increased in blue swimming crab shell (Zn and Cu) and oyster shell (Fe) as a result of increasing the calcination temperature. This study demonstrates the optimum calcination temperature of calcium oxide production from four types of marine wastes that might be benefit for the chemical compound production industry.  相似文献   

4.
The early stages of ceria growth on Rh(111) at high temperature have been investigated by low-energy electron microscopy and photoemission electron microscopy. Ceria was deposited by reactive Ce deposition at substrate temperatures between 700°C and 900°C in an oxygen ambient of 5 × 10−7 Torr. At 700°C, we observe a high nucleation density of 100-nm-sized islands. With elevated temperature, the average island size increases, and the nucleation density decreases. Triangularly shaped islands nucleate preferentially at step edges, with seemingly abrupt interfaces between Ce and Rh. At 900°C, the island edges are still straight, but during growth the islands lose their triangular form. Instead, growth along the substrate step edges becomes favorable, leading to a maze-like morphology. Atomic force microscopy reveals islands of 0.3 to 0.6-nm height, consistent with ceria islands formed by one or two trilayers (O―Ce―O) of ceria. Moreover, the second layer of the islands is also triangularly shaped, with lateral dimensions of 50 nm and similar step heights. IV-LEEM analysis leads to the conclusion that the rhodium surface is covered by a layer of reduced cerium oxide, which is partially overgrown by smaller islands of CeO2.  相似文献   

5.
李良彬 《高分子科学》2014,32(9):1224-1233
In this study, recovery processes of isotactic polypropylene(iPP) melted spherulites at 135 °C after melting at higher temperatures(170 °C–176 °C) were investigated with polarized optical microscopy and Fourier transform infrared spectroscopy. The recovery temperature was fixed to exclude the interference from heterogeneous nuclei. After melting at temperatures between 170 °C and 174 °C, the melted spherulite could recover back to the origin spherulite at low temperatures. Interestingly, a distinct infrared spectrum from iPP melt and crystal was observed in the early stage of recovery process after melting at low temperatures, where only IR bands resulting from short helices with 12 monomers or less can be seen, which indicates that the presence of crystal residues is not the necessary condition for the polymer memory effect. Avrami analysis further indicated that crystallization mainly took place in melted lamellae. After melting at higher temperatures, melted spherulite cannot recover. Based on above findings, it is proposed that the memory effect can be mainly ascribed to melted lamellae, during which crystalline order is lost but conformational order still exists. These conformational ordered segments formed aggregates, which can play as nucleation precursors at low temperatures.  相似文献   

6.
In view of the susceptibility of TiB2 to oxidation, the thermal stability of monolithic TiB2 and of Al2O3-30 vol% TiB2 and Si3N4-20 vol% TiB2 composites was investigated. The temperature at which TiB2 ceramic starts to oxidize is about 400°C, oxidation kinetics being controlled by diffusion up toT≈900°C and in the first stage of the oxidation at 1000°C and 1100°C (up to 800 min and 500 min respectively), and by a linear law at higher temperatures and for longer periods. Weight gains in the Al2O3-TiB2 composite can be detected only at temperatures above ≈700°C and the rate governing step of the oxidation reaction is characterized by a one-dimensional diffusion mechanism atT=700°C andT=800°C and by two-dimensional diffusion at higher temperatures. Concerning the Si3N4-TiB2 composite, three different oxidation behaviours related to the temperature were observed, i.e. up to ≈1000°C the reaction detected regards only the second phase; at ≈1000<T<≈1200°C, the diffusion of O2 or N2 through an oxide layer is proposed as the rate-governing step; atT〉=1200°C, a linear kinetic indicates the formation of a non protective scale.  相似文献   

7.
《Comptes Rendus Chimie》2016,19(10):1326-1336
This review is focused on the use of the water splitting reaction for characterizing oxygen vacancies in supported metal catalysts and more generally to get insight into the high-temperature modifications of metal–support interactions. Three supports widely used in catalysis are considered, namely alumina, silica and ceria. The catalysts were reduced at temperatures TR ranging from 200 to 1000 °C. The reaction with water was carried out at temperatures TOX ranging from 100 to 1000 °C. In every case, the metal (Rh or Pt) was chosen among those which are not oxidizable by water. Extensive investigations of the reactivity of water with unsupported metals and films confirmed this choice. The reaction is then selective for the titration of O vacancies, generally associated with reduced cations of the support. On alumina-supported catalysts, reduction at TR > 600 °C leads to the formation of oxygen vacancies strictly confined to the periphery of metal particles. The amount of hydrogen produced QH is coherent with the peripheral oxygen density. Reduction of silica-supported catalysts at TR > 600 °C generates metal silicides that can be selectively destroyed by water with reformation of silica and metal nanoparticles. Oxygen vacancies are formed on ceria catalysts at 200 °C. These oxygen vacancies are confined to the surface up to 600 °C. At higher temperatures, oxygen vacancies are formed in the bulk: about 50% of CeO2 would be reduced at 900 °C. The amount of H2 produced by reaction with water is thus very high on metal-ceria catalysts. At TR > 900 °C, metal cerides start to form. Remarkably, a significant reactivity of H2O on a Rh/CeO2 catalyst reduced at 850 °C is recorded as of 100 °C. However, the quantitative titration of oxygen vacancies required temperatures TOX > 500 °C. As a rule, the technique of water splitting allows the detection of 1 μmol g−1 of oxygen vacancies, i.e. a few 0.1% of the surface in the case of reducible oxides of 10–20 m2 g−1.  相似文献   

8.
The reaction of CO oxidation on thin metal films of platinum, rhodium, and their alloy and on a heterophase bimetallic Pt/Rh surface that consisted of platinum particles of size 10–20 nm on the surface of rhodium was studied in the region of low reactant pressures (lower than 2 × 10?5 mbar). At low temperatures (T < 200°C), the activity of samples increased in the order Rh > Pt/Rh > Pt-Rh alloy > Pt. Above 200°C, the rate of reaction on the heterophase Pt/Rh surface was almost twice as high as the sum of the rates of reaction on the individual metals; this fact is indicative of a synergistic effect. The nature of this effect is considered.  相似文献   

9.
Adsorption isotherms have been measured on cleaned silver powder from 178 to 339°C at oxygen pressures of 0.226 Pa to 40 kPa using a vacuum ultramicrobalance. Adsorption equilibrium was found at all temperatures and pressures studied. The surface was prepared for the reproducible chemisorption studies using an established method of cyclic outgassing, oxygen adsorption and reduction in carbon monoxide. Seven isotherms were measured that spanned fractional surface coverages from 0.17 to 1.1. The isosteric heat of adsorption q was determined at constant values of θ. After decreasing from 42 to 17.7 kcal mol?1 at the lower coverages, q remains constant at 18.4 ± 0.8 kcal mol?1 from θ of 0.33 to about 0.90 and then decreases to zero at the highest coverages and temperatures. The initial drop in q is attributed to the formation of islands of a two-dimensional surface silver oxide. The constant value of q results then from the completion of the oxide layer and molecular adsorption on and/or through the oxide. The decrease in q to zero at the highest coverages results from repulsions in the adlayer at T ≤ 275°C and absorption into silver at T > 302°C.  相似文献   

10.
Synthetic and natural goethites (0.5–1.5 mg) were heated up to 600°C in alkali-halide disks (400 mg). The thermal transformations occurring at different temperatures are found to depend on the preparation of the disks. For mixtures of alkali-halides and goethite not ground during the preparation of the disks, heating at >200°C resulted in protohematite, which persisted up to 600°C. However, disks which were subjected to repeated grinding—pressing cycles before thermal treatments gave rise to protohematite at >200°C, which on further heating at >300°C was transformed to a transitional iron oxide. In CsI disks, the transitional oxide derived from synthetic goethite can be further transformed to maghemite at 500°C; however, almost no maghemite could be obtained from natural goethite. At 600°C, both the transitional oxide and the maghemite resulting from the synthetic goethite in CsI disks were reduced to magnetite. On the other hand, in KI disks, transitional oxides obtained from both synthetic and natural goethites were reduced to magnetite upon re-pressing and gradual heating of the disks at 600°C. In KI disks, magnetite can be formed only if the reduction temperature is reached gradually, whereas in CsI disks magnetite is formed upon direct heating of the disks to 600°C. The iron oxides referred to above, including the transitional oxides resulting from thermal treatments, were studied by IR absorption spectroscopy.  相似文献   

11.
The influence of lithium oxide-doping on the thermal stability of Co3O4 was studied using DTA, TG, DTG and X-ray diffraction techniques. Pure and doped cobaltic oxide specimens were prepared by thermal decomposition of pure basic cobalt carbonate and the basic carbonate mixed with different proportions of LiOH, in air, at different temperatures between 500 and 1100°C.Pure Co3O4 was found to start partial decomposition when heated in air at 830°C yielding the CoO phase. The complete decomposition was effected by heating at 1000°C.Doping of Co3O4 with different proportions of Li2O was found to much increase its thermal stability. The temperatures at which the doped oxide samples started to undergo decomposition were increased to 865, 910 and 1050°C for 0.375, 0.75 and 3% Li2O-doped solids, respectively. The DTA revealed that the 1.5% Li2O-doped cobaltic oxide did not undergo any thermal decomposition till 1080°C. The X-ray investigation showed that the prolonged heating of 1.5 and 3% Li2O-doped solids at 1100°C for 36 h effected only a partial decomposition of Co3O4 into CoO. Heating of these solids at temperatures varying between 900 and 1100°C led also to the formation of a new lithium oxide cobaltic oxide phase, the composition of which has not yet been identified.The role of Li2O in increasing the thermal stability of Co3O4 was attributed to the substitution of some of its cobalt ions by Li+ ions, according to Verwey and De Boer's mechanism, leading to the transformation of some of the Co2+ into Co3+ ions thus increasing the oxidation state of the cobaltic oxide lattice.  相似文献   

12.
The polyaniline-PbO composites of various mass fractions were prepared by in situ polymerisation. The prepared samples were characterised by FTIR, and the dominant peaks confirmed the formation of polyaniline-PbO composites. The SEM study shows a granular agglomerated morphology, and increases with an increase in the lead oxide mass % in polyaniline. Direct current (DC) conductivity (σ DC) was studied as a function of temperature (T). From these studies, it was found that conductivity increased at higher temperatures due to the polarons hopping from one localised state to another. DSC studies reveal, the decrease in peak temperature from 273°C (pure PANI) to 169.2°C, 193.5°C, 218.4°C, 235.2°C, and 224.2°C, respectively for the various mass fractions (10 %, 30 %, 20 %, 40 %, and 50 %) of polyaniline-PbO composites.  相似文献   

13.
The mechanism of vaporization and atomization of U in a graphite tube electrothermal vaporizer was studied using graphite furnace atomic absorption spectrometry (GFAAS) and electrothermal vaporization inductively coupled plasma mass spectrometry (ETV-ICP-MS). Graphite furnace AAS studies indicate U atoms are formed at temperatures above 2400°C. Using ETV-ICP-MS, an appearance temperature of 1100°C was obtained indicating that some U vaporizes as U oxide. Although U carbides form at temperatures above 2000°C, ETV-ICP-MS studies show that they do not vaporize until 2600°C. In the temperature range between 2200°C and 2600°C, U atoms in GFAAS are likely formed by thermal dissociation of U oxide, whereas at higher temperatures, U atoms are formed via thermal dissociation of U carbide.The origin of U signal suppression in ETV-ICP-MS by NaCl was also investigated. At temperatures above 2000°C, signal suppression may be caused by the accelerated rate of formation of carbide species while at temperatures below 2000°C, the presence of NaCl may cause intercalation of the U in the graphite layers resulting in partial retention of U during the vaporization step. The use of 0.3% freon-23 (CHF3) mixed with the argon carrier gas was effective in preventing the intercalation of U in graphite and U carbide formation at 2700°C.  相似文献   

14.
We studied the effects of temperature on the sulfate permease of Penicillium chrysogenum (PC) (a mesophile with a growth temperature range of 4-35°C) and Penicillium duponti (PD) (a thermophile with a growth temperature range of 27-58°C). Arrhenius plots of sulfate permease activity from mycelia grown at 50°C (PD), 30°C (PD and PC) or 8°C (PC) indicate that at temperatures below the transition point there is little difference in the activation energy of sulfate permease in mycelia from PD grown at 50°C or 30°C or PC grown at 30°C or 8°C; however, the temperature of the transition point for the permease from each set of mycelia assayed reflects the optimum growth temperature of the fungal source. Transitions occur at 15 °C for mycelia from PC and 35 ° C for PD mycelia. Kinetic measurements indicate that the Km of sulfate permease in PC cells grown at a variety of temperatures is essentially the same at various temperatures. As an example, the Km of 8°C or 30°C grown PC is about 55 μM at 25°C and 45 μM at 8°C. Vmax measurements reflect growth conditions such as temperature and growth stage. p]Lipid composition of the mycelia dramatically reflect growth temperatures. Double bond index values vary from 1.94 for PC grown at 8°C to 0.81 for PD grown at 50°C. The percentage of total fatty acid represented by linolenic acid varies from 45% in 8°C grown PC to 4.2% or less in 30°C grown PC. No linolenic is found in mycelia from PD.  相似文献   

15.
《Mendeleev Communications》2023,33(4):574-576
The ignition temperatures and effective activation energies of the ignition limits of mixtures (40–70% H2 + 60–30% CH4)stoich + air over Rh were experimentally determined at a pressure of 1 atm in the temperature range 20–300 °C. Over an ignition-treated Rh surface, the ignition temperature of a mixture of 70% H2 + 30% methane + air is 62 °C. This indicates the potential of using Rh to markedly lower the ignition temperature of fuels based on hydrogen–methane mixtures.  相似文献   

16.
An in situ comparative study of the reduction of Co-containing catalysts for the Fischer–Tropsch process in hydrogen and supercritical (SC) isopropanol is performed by ferromagnetic resonance (FMR) spectroscopy. According to the FMR data, the reduction of cobalt-containing oxide particles to metal in hydrogen starts at temperatures of ~360°C, which is substantially lower than a temperature of the formation of metal particles of the active phase according to powder X-ray diffraction and differential thermogravimetry data (Т ~ 450°C). In SC isopropanol, the reduction to Co metal occurs at lower temperatures (T ~ 245°C) as compared with the reduction temperature for these catalysts in hydrogen. It is shown that the reduction in SC isopropanol can lead to the formation of superparamagnetic Co nanoparticles with a narrow particle size distribution.  相似文献   

17.
The crystal structure of (η5-C5Me5)(η3-MeHCCHCHMe)RhCl at ?120°C was determined (R = 0.041 for 1790 reflections). The molecule has approximate mirror symmetry. The cyclopentadienyl ring is bent by 6.8°, acquiring an envelope-like conformation, and its bonding with Rh is of a partially-localized η41 type with RhC(“π-diolefin”) of 2.206–2.235 Å and RhC(“σ-bonded”) of 2.151 Å. The syn-arrangement of the Me groups in the π-allyl ligand, assigned by NMR spectra, is confirmed.  相似文献   

18.
Two different hybrid materials composed of gold nanoparticles (AuNPs) supported on either commercial niobium oxide HY 340 or mesoporous niobium oxide catalyzed the Friedel-Crafts alkylation of anisole by benzyl chloride. Excitation of the surface plasmon of the supported AuNPs allowed the reaction to occur at lower temperatures by acting as an alternative heat source. The localized heating produced via plasmon excitation permitted the acid catalyzed reaction to occur - at the Lewis acid sites on the Nb2O5 support - at 80 °C while thermal-dark reactions using a conventional heat source, required temperatures of 120 °C or higher. The catalytic activity of the tested hybrid materials decreased with storage time. However, the deactivation showed to be reversible upon lyophilisation indicating that the nature of the deactivation could be due to water adsorption.  相似文献   

19.
Demand for high-density press and sinter components is increasing day by day. Of the different ways to improve the sinter density, the addition of nanopowder to the conventional micrometer-sized metal powder is an effective solution. The present investigation is aimed at studying the surface chemistry of iron nanopowder coated with graphitic carbon, which is intended to be mixed with the conventional iron powder. For this purpose, iron nanopowder in the size range of 30 nm to submicron (less than 1 micron) was investigated using thermogravimetry at different temperatures: 400°C, 600°C, 800°C, 1000°C, and 1350°C. The X-ray photoelectron spectroscopy (XPS), Auger electron spectroscopy (AES), and high-resolution scanning electron microscopy (HR-SEM) were used for characterizing the powder as well as samples sintered at different temperatures. The presence of iron, oxygen, carbon, chromium, and zinc were observed on the surface of the nanopowder. Iron was present in oxide state, although a small metallic iron peak at 707 eV was also observed in the XPS spectra obtained from the surface indicating the oxide scale to be maximum of about 5 nm in thickness. For the sample treated at 600°C, presence of manganese was observed on the surface. Thermogravimetry results showed a two-step mass loss with a total mass loss of 4 wt.% when heated to 1350°C where the first step corresponds to the surface oxide reduction.  相似文献   

20.
Carbon steels (CSs) were anodized in an ethylene glycol solution containing 3 vol.% H2O and 0.1 m NH4F to coat with nanotube arrays film. The as anodized nanotube arrays film were annealed in argon atmosphere at various temperatures ranging from 250 to 550 °C for 4 h. The morphology and crystal phases of the film developed after annealing processes were examined using field emission scanning electron microscopy, X‐ray diffraction. Morphology transforms from nanobube arrays to nanotube bundles at 250 °C, to nanobube bundles with nanoflakes at 350 and 450 °C, to nanotube bundles with nanobelts at 550 °C. Amorphous transformed completely into maghemite at 350 °C and hematite with minor magnetite at 450 and 550 °C. Diffuse reflectance ultraviolet and visible spectra revealed iron oxide nanotube film annealed at 350 °C, or higher than 350 °C behaved tremendous absorbance ability in visible spectra range. Mott–Schottky analysis and linear scan voltammetry were performed in 1 m NaOH to show that iron oxide nanotube film annealed at 450 °C exhibited best charge carrier transfer ability upon illumination and superior photoelectrochemical properties compared with the films annealed at other temperatures. The film annealed at 450 °C displayed the photocurrent density of 0.13 mA cm?2 at 0.2 VAg/AgCl, but the film annealed at other temperatures with the photocurrent densities of lower than 0.05 mA cm?2 at 0.2 VAg/AgCl. The morphology and phase transform of iron oxide nanotube film at different annealing temperature results in the change of their photoelectrochemical properties. Copyright © 2016 John Wiley & Sons, Ltd.  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