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.
Two-phase layered scales comprising CoO and Co 3O4 formed on cobalt during oxidation at 600°, 700°, and 800°C and at oxygen partial pressures in the range 0.001–1 atm. The kinetics, which were obtained by thermogravimetric analysis, obeyed a parabolic rate law after an initial, non-parabolic stage of oxidation. The monoxide consisted of relatively large grains (10 ) and the spinel comprised small grains (3 ) for all conditions of oxidation. Grain boundary diffusion of cations played a significant role in the growth of the spinel layer. Thermogravimetric data and the steady-state ratio of the oxide layer thicknesses were employed to calculate the rates of thickening of the individual oxide layers and the rate of oxidation of CoO to Co3O4.  相似文献   

2.
Thermal oxide films grown on electropolished aluminum specimens have been investigated by transmission electron microscopy of stripped oxide films and ultramicrotomed sections. Particular attention has been focused on the nucleation sites -Al 2 O 3 crystals and the relationship of such sites to surface features on the electropolished aluminum surface. It is evident that easy paths for the diffusion of oxygen, or the nucleation sites of -Al 2 O 3 crystals, are not distributed randomly over the electropolished aluminum surface, but form preferentially in the amorphous oxide layer grown over preexisting metal ridges. Thus, the diffusion of molecular oxygen through cracks in the amorphous oxide layer represents the most realistic and acceptable basis for explaining the local growth of the -Al 2 O 3 crystals in thermal oxide films on aluminum, although the cracks have not yet been observed directly.Present address: Alcan International, Ltd., Banbury Laboratories, Banbury, Oxford, OX16 7SP, United Kingdom.  相似文献   

3.
The oxidation behavior of -Ni3Al and -Ni3Al(Si) (Ni75Al20Si5) intermetallic compounds was studied at 1073 K and oxygen partial pressures of 59×10–6 atm, 1.2×10–14 atm, and 1.2×10–19 atm by means of a manometric apparatus, and Rhines packs of NiO/Ni and FeO/Fe, respectively. Oxidation kinetics were determined either by recording weight gains or by measuring the internal-oxidation-zone depths. The structures and morphologies of oxides were also studied. Relatively low oxidation rates occurred for both compounds when oxidized in the manometric apparatus, while fast internal oxidation was observed for both compounds in the NiO/Ni pack. The fastest oxidation occurred in the -Ni3Al compound in the FeO/Fe pack. A healing Al2O3 layer was formed on the surface of -Ni3Al(Si) compound in the FeO/Fe pack, indicating a synergistic effect between the solute elements in the compound at the oxygen pressures corresponding to the dissociation of FeO. The oxidation rate was found to depend on the volume expansion associated with solute-atom oxidation.  相似文献   

4.
A critique of internal oxidation in alloys during the post-wagner era   总被引:1,自引:0,他引:1  
Wagner's classical treatment of internal oxidation (generic name allowing for reaction with oxygen, nitrogen, carbon or sulfur) assumed ideal conditions such as uninhibited dissolution of the gas, formation of spherical particles, diffusion of the oxidant in the solvent as the rate-controlling step, equilibrium conditions, etc. However, during the 45 years since his treatment, many observations have been made to complicate the idealized situation suggested by Wagner. This paper examines the most important modifications with respect to Wagner's original analysis. The following items are discussed. (a) The role of solute concentration: The parabolic kinetics are much higher than expected for Ni–Al alloys due to rapid interfacial diffusion of oxygen along the interfaces between cylindrical rods of Al2O3 (perpendicular to the surface) and the matrix. (b) Precipitate morphology: Spherical precipitates seem almost to be the exception. A wide variety of forms have been observed, including Widmanstätten platelets, cylindrical rods, hexagonal plates, dendritic or fishbone products, etc. The competition between nucleation and growth is useful to explain the observed structures. (c) Intergranular internal oxidation: Rapid oxygen diffusion in grain boundaries may lead to a wide variety of intergranular-precipitate structures. (d) Internal-oxide bands: Wavy, approximately parallel bands form at a finite distance beneath the surface in certain alloys having very reactive solutes, e.g., Ag–Mg. It is postulated that high stresses generated by precipitation play a major role. (e) Surface nodules of pure solvent metal: High stresses generated during precipitation cause extrusion of solute through dislocation pipes, leading to extensive nodule formation on either grain boundaries or on the grains (or both), depending on the alloy and oxidizing conditions. (f) Nonstoichiometric precipitates: Either hypo- or hyperstoichiometric particles can form as very small clusters in certain alloys (Ag–Al). The nature of precursors and changes in stoichiometry during reaction are discussed. (g) Trapping of oxidant: Diffusion of the oxidant may be slowed appreciably by trapping with the solute, although no precipitates need to form. Lower-than-expected kinetics (based on normal diffusivities of the oxidant) result. (h) High-solubility-product precipitates: Concentration profiles of solute, oxidant and precipitate are quite different than those expected for low-solubility-product precipitates as considered by Wagner. In particular, a variable mole fraction of precipitate exists, and further precipitation occurs in the reaction zone after the front has passed by. Linear kinetics have been observed for some Nb-base alloys at very high temperatures and low oxygen pressures. The rate-controlling step is the arrival of oxygen at the surface and not oxygen diffusion in the metal. (i) Dual oxidants: Two gases may diffuse·simultaneously and each forms its own product with the solute. The thermodynamically most-stable compound forms near the surface, and the less-stable compound deeper in the alloy. The less-stable compound is subsequently converted to the more-stable compound with a concomitant release of the second oxidant. Although numerous examples have been reported of systems which do not behave as predicted by Wagner, his theory still remains as the cornerstone of our understanding and is still the starting point for virtually every study in internal oxidation.  相似文献   

5.
Nickel-base alloys, such as Astroloy, used for aeronautical turbine disks, are sensitive to time-dependent cracking in environments containing oxygen. The mosaic structure of the alloy consisting islands (200 nm average size) surrounded by the -phase (100 nm thick) induces complex oxidation phenomena. Various analytical approaches allow the delineation of all the steps from segregation to oxidation occurring on the surface of such a duplex structure. The protection of Astroloy by its outer oxide layer against oxygen penetration was studied also, using alternative 16O2 then 18O2 oxidation. In association with STEM studies, it is shown that the outer oxide scale is not a real barrier against oxygen penetration and that inner precipitation of chronium (+ aluminium and titanium)-enriched oxides, takes place especially in the structure.  相似文献   

6.
The effect of a SiO 2 layer, partially removed by wet chemical processes or thermally grown, on the mechanism of nucleation and growth of Cu 3 Si from the reaction between a Si(100) wafer and gaseous copper chloride was studied. For oxide layer thicknesses less than 2 nm, the number of Cu 3 Si nuclei and the area of silicon reacting with CuCl per unit area are inversely proportional to thickness, whereas the size of the more numerous nuclei increases from 4–10 m. In this case, the nuclei are octahedrally shaped. The superficial erosion of Cu 3 Si nuclei and the formation of pits have been explained by the reaction between Cu 3 Si and CuCl. For thicker layers, 10e40nm, the Cu 3 Si nuclei are deformed, and the reaction is strongly inhibited. A SiO 2 layer of 45 nm prevented active-site formation.  相似文献   

7.
The oxidation behavior of an Fe-27%Cr-4%Al alloy and similar alloys containing 0.023% and 0.82% Y in 1 atm oxygen at 1200°C has been examined. The oxide formed on the yttrium-free alloy develops a highly convoluted configuration, apparently resulting from lateral growth of the oxide. The latter leads to oxide detachment from the alloy at temperature and extensive spalling during cooling. It is postulated that lateral growth results from the formation of oxide within the existing oxide layer by reaction between oxygen diffusing inward down the oxide grain boundaries and aluminum diffusing outward through the bulk oxide. Additions of yttrium to the alloy apparently prevent the formation of oxide within the oxide layer, the oxide-forming reaction occurring as the alloy-oxide interface. Thus lateral growth is prevented and spalling during cooling does not occur. Secondary advantages conferred by the addition of 0.82% Y to the alloy are the prevention of void formation at the alloy-oxide interface, the avoidance of alloy grain growth during oxidation, and the creation of an oxide keying or pegging effect.  相似文献   

8.
The reaction of oxygen at low pressures with silicon layers on tungsten ribbons was studied. An abrupt transition was observed between a condition of passivation, in which a thin film of SiO2 formed at low temperatures, and a steady-state combustion condition at high temperatures. The latter state is characterized by the formation of volatile SiO. The boundary between these two states has been defined in terms of the pressure-temperature relation. Oxygen consumption in the combustion state is represented by first-order reaction kinetics with an activation energy of 13 ± 1 kcal/mole. The stability of the two states has been defined by a thermodynamic analysis of the SiO2 layer stability. The oxygen consumption dependence on temperature has been described by a kinetic model which involves a consideration of the various elementary steps in the reaction.  相似文献   

9.
Structurization phenomena in a mixed solvent were found to determine, by affecting the solvation degree of an anion, its adsorptivity at the electrode and, consequently, the surface concentration of the nonaqueous solvent. The formation of activated Cd2+ complexes with iodide anions in an adsorbed layer is usually accompanied by accelerating the electrode reaction. The desalting effect of the mixed solvent in the zones of its structural stabilization is mainly manifested as the enhanced adsorptivity of iodide anions and the correspondingly facilitated discharge of metal ions. The highest discharge rate was found at 2 0.9 when the structure of EtOH is ordered with monomeric water molecules. The highest surface concentration of EtOH and the lowest rate of cadmium electroreduction correspond to a structurally disordered mixture ( 2 0.4).  相似文献   

10.
Fe-30Mo alloys containing up to 9.1 wt% Al were sulfidized at 0.01 atm sulfur vapor over the temperature range of 700–900°C. The sulfidation kinetics followed the parabolic rate law for all alloys at all temperatures. For alloys containing small and intermediate amounts of Al (<4.8 wt.%), a duplex sulfide scale formed. The outer layers of the scales were found to be relatively compact FeS in all cases; whereas the inner layers were composed of the layered compound MoS 2 (intercalated with iron), the Chevrel compound Fe x Mo 6 S 8,a spinel double sulfide Al x Mo 2 S 4,depending on the Al content of the alloy and the sulfidation temperature. Extremely thin scales were found on the alloys with higher Al contents. Accordingly, extremely slow sulfidation rates were observed—even slower than the sulfidation rate of pure Mo. The transition of the sulfidation kinetics from a high-rate active mode to a low-rate passive mode requires both a critical Al content in the alloy and a critical Mo content. Because of the two-phase nature of the alloys, the latter requirement implies a critical volume fraction of the intermetallic second-phase in the alloy, which has been known as the multiphase effect. Interestingly, the multiphase effect in these alloys was also a function of the Al content in the alloys.  相似文献   

11.
The oxidation behavior is very different for an aluminide with a wide homogeneity range such as -NiAl than for a line compound such as NbAl 3.Oxidation of -NiAl at temperatures 1273 K leads to a slow-growing -alumina layer. The metal phase beneath the scale remains as -NiAl; however, cavity formation is observed. The cavity formation may be favored by sulphur surface segregation. Oxidation of NbAl 3 at temperatures 1273 K initially leads to -Al 2O3,but the Al depletion causes the formation of Nb 2 Al beneath the oxide layer. Cracking of the Al 2O3 layer opens Nb 2 Al to the atmosphere, which oxidizes rapidly to Nb 2O5 and NbAlO4.After consumption of the Nb 2 Al, a layer of Al 2O3 formed again on the NbAl 3 phase, but failure of the alumina and the fast growth of the other oxides occur as a repeated process. Thus, NbAl 3 exhibited rapid linear oxidation kinetics. Multiphase alloys in the system Nb-Ni-Al generally behave better than NbAl 3,and the low oxidation rates of -NiAl can be approached. In the temperature range below 1273 K, with a maximum at 1000 K, both NiAl and NbAl 3 show the pest phenomenon, an intergranular disintegration. Preceding the disintegration, oxygen diffuses into the grain boundaries of the material and Al 2O3 is formed at the grain boundaries, beginning from the surface region. NiAl is susceptible only in a very limited range of oxygen pressures and temperatures, whereas NbAl 3 is much more susceptible.  相似文献   

12.
A series of oxide-dispersed-NiAl alloys were oxidized in order to explore the effect of various cation dopants on the - phase transformation in the Al2O3 scale and the effect of phase composition on the scale microstructure. Larger ions such as Y, Zr, La, and Hf appeared to slow the- to-Al2O3 phase transformation, while a smaller ion, Ti, appeared to accelerate the transformation.  相似文献   

13.
A marker study of nickel oxidation in SO2 at 600°C has proved that after the reaction metallic markers (Au, Pt, W) are covered with a sulfide rim. This effect is not observed on quartz markers. The metallic markers make the intermediate NiO layer adjacent to them increase, whereas the quartz marker makes this layer disappear.  相似文献   

14.
Preoxidized chromium specimens have been high vacuum annealed at 1200° and 1300°C to produce densified Cr 2O3 scales. These specimens have been reoxidized at the same temperatures at 10–6 atm O2. The initial reoxidation is linear with time and is concluded to reflect a volume diffusion controlled transport through the densified scale. The corresponding parabolic rate constant (w2 = kpt)is given by kp=1.4 · 10–2 exp(–235,000/RT)(gram of O) 2/cm4 sec. It is tentatively concluded that outward chromium diffusion predominates in an inner layer of the Cr2O3 scales and inward oxygen diffusion in an outer layer. Under the experimental conditions it has not been possible to maintain growth of the Cr2O3 scales controlled by volume diffusion. The new oxide layer consists of fine crystallites; the oxide grows at grain boundaries within the scales. This causes sideways growth of the scale, breakdown of the originally densified layer, and an increased rate of reaction.  相似文献   

15.
The equilibrium oxygen potentials of the two-phase equilibria Fe3O4/Fe1–xS, Co1–xS/Co3S4, Co3S4/CoS2, Co1–xS/CoO, and CoO/CoSO4 were measured as a function of temperature. A solid-state emf technique using calcia-stabilized zirconia (CSZ) solid electrolyte was used. These equilibria were studied atP SO21 atm; the equilibrium Co1–xS/CoO was also studied atP SO20.1 atm. Two emf cell designs were used for the measurements atP SO21 atm andP SO21 and 0.1 atm, respectively. The homogeneity range of FeS in equilibrium with Fe3O4 and that of Co1–xS in equilibrium with CoO at 1073 K andP SO21 atm were measured by electron microprobe analysis.  相似文献   

16.
A model is proposed for the oxidation of metals including both bulk crystal lattice diffusion and short-circuit diffusion paths. Assuming local equilibrium between point defects in the bulk and in grain boundaries, we obtain an effective parabolic rate constant keff=kbulk [1+ exp-(H/RT],where H is the enthalpy for the reaction between point defects in the bulk and in short circuits, is the fraction of the short-circuit area, and T is temperature. In the case of the high-temperature oxidation of nickel coated with a thin film of vanadium pentoxide, this model yields to a rate law with a critical oxygen pressure P c. The nickel vanadium oxide located in the short circuits of the growing oxide NiO is liquid below Pc, leading to a rapid short-circuit diffusion. Above this value, a vanadate precipitates and acts as diffusion blocks for migrating Ni ions.  相似文献   

17.
The growth of Al2O3 scales on -NiAl was studied at 1500°C. Oxidation rates, diffusion mechanisms, and microstructures were examined in order to achieve a complete understanding of the scale development. Variation of the Al content within the phase field had little effect on the oxidation behavior. Ionimplanted yttrium (2×1016/cm2) was observed to provide a short-term improvement in scale adhesion but little long-term effect. When doped with Y or Zr, the first 1 m of -Al2O3 was observed to grow mainly by an inward oxygen growth mechanism. At longer times when the implant was ineffective, microstructural observations indicate a mixed-growth mode.  相似文献   

18.
An electrochemical method based on analyzing the potential-current curves, plotted at various applied PO 2 for an alumina scale developed at 1100°C on a -NiAl alloy, was used to determine transport parameters in the scale. The variation of Vo vs PO 2 indicates that the scale consists of an inside zone characterized by an ionic-transport number ti0.4 and an outside part with ti0.1. This leads to a value of the oxygen pressure at the NiAl/Al2O3 interface 2×10–27 atm, not far from the equilibrium value. The variation of the oxygen chemical potential inside the scale was determined, showing a steep variation of µO in the middle of the scale. The calculated oxidation constant is close to the experimental one, which would indicate that mainly charged species are responsible for the alumina scale growth.  相似文献   

19.
Theoretical predictions computed on the basis of the quasiequilibrium treatment of gas-solid reactions are compared with existing experimental data on the rate of volatilization (erosion) of solid tungsten by reaction with gaseous O2 at high temperature ( 1300° T 3600° K) and low pressure (4.5 × 10–7 11.5 Torr). The only unknown parameter in the analysis is the equilibrium probability, , defined as the fraction of the impinging O2 molecules that attain thermochemical equilibrium at the tungsten surface rather than undergoing nonreactive scattering (e.g., reflection). An approximate expression for is estimated by a straightforward empirical procedure that is consistent with the quasiequilibrium treatment. The theoretical results based on this expression for T because appears to be an exponential function ofT; (b) In the intermediate region, the formation of volatile oxides decreases sharply with increasingT because atomic oxygen becomes the thermodynamically favored reaction product, thereby causing W to decrease with increasingT; (c) In the highest region, W again increases withT as a result of the formation of WO and the sublimation of W.This work was supported by the Joint Services Electronics Program [Contract DA28-043-AMC-02536(E)] and by NASA [Grant NGR-22-009-091].  相似文献   

20.
It is shown that the strongest deceleration of the electrodeposition of cadmium in inhibited media is observed in the domain of loosening the structure of a mixed solvent (0.45 x 2 0.25) when the molecular adsorption of organic solvent on Cd is maximum. The process is maximally facilitated in the domain of ordering the acetonitrile's (AN) structure (0.9 x 2 0.75), where perchlorate anions dominate at the cathode surface. Crown-ester exhibits maximum inhibition effect under the conditions of preferential adsorption of its molecules, i.e., in the domain of the prevalence of the nonaqueous component in the mixture (x 2 0.95).  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