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 31 毫秒
1.
The oxidation of two Co–Nb alloys containing 15 and 30 wt.% Nb has been studied at 600–800° C in H2–CO2 mixtures providing an oxygen pressure of 10–24 atm at 600°C and 10–20 atm at 700 and 800°C, below the dissociation pressure of cobalt oxide. At 600 and 700°C both alloys showed only a region of internal oxidation composed, of a mixture of alpha cobalt and of niobium oxides (NbO2 and Nb2O5) and at 700°C also the double oxide CoNb2O6, which formed from the Nb-rich Co3Nb phase. No Nb-depleted layer formed in the alloy at the interface with the region of internal oxidation at these temperatures. Upon oxidation at 800°C a transition between internal and external oxidation of niobium was observed, especially for Co–30Nb. This corrosion mode is associated with the development of a single-phase, Nb-depleted region at the surface of the alloy. The corrosion mechanism of these alloys is examined with special reference to the effect of the low solubility of niobium in cobalt and to the relation between the microstructures of the alloys and of the scales.  相似文献   

2.
A study has been made on the influence of surface preparation involving abrasion, electropolishing, and vacuum annealing on the physical nature of the (100) nickel crystal face and on the morphological development of the oxide film. The structure of the initial surface was dependent upon the method of surface preparation, but electropolished specimens subjected to conventional or high vacuum anneals at 10–6 and 10–10 Torr, respectively, and temperatures of 800 and 700° C were structurally identical. Thermal faceting of surfaces prepared to correspond to the (100) plane was negligible. Oxidation at 500° C and over the oxygen pressure range 10–5–400 Torr led to formation of polycrystalline oxide crystallites randomly distributed over the surface during early stage exposures; the crystallites were only a few hundred Angstroms in size upon coverage of the surface by a nickel oxide film. These films were relatively uniform for thicknesses up to 2000 Å. Crystallite growth processes led to three major epitaxial relationships between oxide and metal: (100) NiO , (111) NiO , and (211) NiO (100) Ni .This work forms part of a research program sponsored by the National Research Council of Canada.  相似文献   

3.
The corrosion of M–Nb alloys based on iron, cobalt, and nickel and containing 15 and 30 wt% Nb has been studied at 600–800°C under low oxygen pressures (10–24 atm at 600°C and 10–20 atm at 700–800°C). Except for the Co–Nb and Ni–Nb alloys corroded at 800°C, which formed external scales of niobium oxides, corrosion under low O2 pressures produced an internal oxidation of niobium. This attack was much faster than expected on the basis of the classical theory. Furthermore, the distribution of the internal oxide in the alloys containing two metal phases was very close to that of the Nb-rich phase in the original alloys. These kinetic, microstructural, and thermodynamic aspects are examined by taking into account the effects of the limited solubility of niobium in the various base metals and of the two-phase nature of the alloys.  相似文献   

4.
The oxidation of two Ni–Nb alloys containing 15 and 30 wt.% Nb has been studied at 600–800° C in H2–CO2 mixtures providing an oxygen pressure of 10–24 atm at 600° C and 10–20 atm O2 at 700 and 800° C, these pressures being less than the dissociation pressure of nickel oxide. The scales formed on both alloys at 600 and 700° C show only a region of internal oxidation composed of a mixture of alpha nickel and niobium oxides (Nb2O5 or/and NbO2), which formed from both the metal phases present, i.e., Ni8Nb and Ni3Nb. Only small, or even no, Nb depletion was observed in the alloys close to the interface with the zone of internal oxidation at these temperatures. On the contrary, samples of both alloys corroded at 800° C produced a continuous external scale of niobium oxides without internal oxidation. The corrosion mechanism of these alloys is examined with special reference to the effect of the low solubility of niobium in nickel.  相似文献   

5.
Four experimental oxide dispersion strengthened (ODS)Fe-(13–14 at. %)Cr ferritic alloys were exposed for up to 10,000 hr at 700–1100 °C in air and in air with 10vol.% water vapor. Their performance has been compared to other commercial ODS and stainless steel alloys. At 700–800°C, the reaction rates in air were very low for all of the ODS Fe–Cr alloys compared to stainless steels. At 900°C, a Y2O3 dispersion showed a distinct benefit in improving oxidation resistance compared to an Al2O3 dispersion or no addition in the stainless steels. However, for the Fe-13 %Cr alloy, breakaway oxidation occurred after 7,000 hr at 900°C in air. Exposures in 10 % water vapor at 800 and 900°C and in air at 1000 and 1100°C showed increased attack for this class of alloys. Because of the relatively low Cr reservoirs in these alloys, their maximum operating temperature in air will be below 900°C.  相似文献   

6.
The Oxidation of an electro-deposited nanocrystalline Cu (nc Cu) and a conventional coarse-grain Cu (cg Cu) was investigated at 30–800 ° C under 1 atm of oxygen. Both Cu samples formed external scales of copper oxide (Cu2O+CuO). At the lower temperature (30–300 °C) the very slow oxidation rates of both the nc and cg Cu might be attributed to the formation of a protective Cu2O surface layer. However at the higher temperature (300–700 °C), oxidation rates of the nc Cu were obviously faster than those of the cg Cu, which was attributed to faster diffusion of various species along grain boundaries both in the metal and in the scale. In particular, the scale grew faster on the nc Cu by means of not only rapid external oxidation as a result of outward diffusion of Cu-ions but also a significant contribution from inward diffusion of oxygen along the grain boundaries in the scales. Compared with the cg Cu, dissolved O2 in the nc Cu may have a certain effect on the faster oxidation of the nc Cu. Above 700°C, the difference seemed to disappear as a result of the ineffectiveness of grain-boundary diffusion.  相似文献   

7.
The oxidation of two Fe–Nb alloys containing 15 and 30 wt.% Nb has been studied at 600–800°C under low oxygen pressures, similar to those prevailing in environments of the coal-gasification type. The reaction produced only an internal oxidation of niobium to form two niobium oxides (NbO2 and Nb2O5) and in some cases a double Fe–Nb oxide. The kinetics of this reaction were very slow at 600°C but rather fast at 700 and 800°C. A peculiar feature of the internal oxidation of these alloys is that the distribution of the internal oxides follows closely that of the Nb-rich phase in the original two-phase alloys. This behavior, as well as the lack of formation of external scales of niobium oxides, is mainly a result of the limited solubility of niobium in iron and of the consequent presence of two metal phases in the alloys.  相似文献   

8.
Partly oxidized copper films were annealed in a controlled vacuum of 10–7 Pa at a temperature of 450° C. The changes discussed below were observed in situ with a specially designed high-resolution transmission electron microscope. The thin, (100)-oriented, single-crystal films of copper had been oxidized immediately prior to the annealing studies at the same temperature and at an oxygen partial pressure of 7×10 –1 Pa, until the desired fraction of the copper film was converted to oxide. It was observed that the oxide disappeared during annealing as long as some copper was left unoxidized. The disappearance of the oxide is explained as being due to dissociation of the oxide at the oxide-metal interface followed by diffusion of oxygen into the metal and desorption of oxygen from the surface of the unoxidized copper. The rate of disappearance of the oxide was found to be proportional to the surface area of unoxidized copper, i.e., the desorption was found to be the rate — limiting step. In the case of heavily oxidized films (>50%), holes were observed to develop in the oxide near the oxide-metal interface after an annealing period of 2–3 hr. Upon resumption of the oxidation, these holes first disappeared, and the normal oxidation behavior was then resumed. The formation of holes may be explained by vacancy clustering. When completely oxidized films were annealed, recrystallization of the oxide was observed.This work was performed at the Ames Research Center and funded by NASA Grants Nos. NCA2-OP390-403 and NSG-2025.  相似文献   

9.
Scales growing on platinum 22 wt.% aluminum specimens at 1100°C were investigated by the following electrical techniques: open-circuit potential and short-circuit current as a function of time and oxygen pressure; current-voltage relationship at several oxygen pressures; transients in the short-circuit current on rapidly changing the oxygen pressure; changes in the growth rate with applied potential. The growth characteristics of the scale were also investigated using kinetic and metallographic techniques. The apparent parabolic rate constant of oxidation decreases in the first hours of oxidation and stabilizes after 20–30 hr at 2 × 10–13 g2·cm–4·sec–1. Transport in the scale is by lattice and grain-boundary diffusion and the changing kinetics are a consequence of changes in the oxide grain size. It was also determined that the electrical characteristics of the scale were affected by the hydrogen-water or carbon monoxide-carbon dioxide atmospheres used to obtain low oxygen pressures. Due to these complexities and to the difficulty in making a satisfactory oxide-gas electrode only qualitative observations could be made from the electrical measurements. These measurements are compatible with the scale having constant ionic conductivity in the pressure range 1 to 10–15 aim and predominantly electronic conductivity at lower oxygen pressures. The rate of oxidation can be accelerated or retarded by passing a current through the scale.  相似文献   

10.
The oxidation of a 2 1 4 Cr-1Mo steel in dry flowing oxygen has been studied in the temperature range 550–700°C for periods of up to 100 hr. A detailed low-resolution microstructural investigation revealed a layered oxide consisting of a very fine-grained and finely pored innermost layer of doped spinel, a central columnar-grained relatively coarsely pored layer of magnetite, and an outer fine-grained hematite layer with fine pores and covered with whiskers of -Fe2O3. This structure is compared with previous results on Fe and model Fe-Cr alloys, as are the kinetics of the oxidation reaction.  相似文献   

11.
The oxidation of copper studied by electron scattering techniques   总被引:1,自引:0,他引:1  
The oxidation of thin single-crystal (100) and (111) films of copper at pressures of around 10–5 Torr and temperatures of 400 to 700°C has been observed by medium-energy electron diffraction, scanning electron microscopy, scanning and conventional microscopy using diffracted electron beams in the reflection mode, transmission electron microscopy, and transmission electron diffraction. Epitaxed nuclei of oxide are observed to grow into very thin single-crystal plates, using oxygen previously trapped in the copper film. There is considerable diffusion of the copper film. There is considerable diffusion of the copper away from the oxide. Between the oxide crystallites the copper surfaces appear to be unoxidized. A mottled contrast of the diffracted beam images of the copper surface is shown to result from many-beam dynamical diffraction effects.  相似文献   

12.
The corrosion behavior of five Fe-Al binary alloys containing up to 40 at. % Al was studied over the temperature range of 700–900°C in a H2/H2S/H2O mixture with varying sulfur partial pressures of 10–7–10–5 atm. and oxygen partial pressures of 10–24–10–2° atm. The corrosion kinetics followed the parabolic rate law in all cases, regardless of temperature and alloy composition. The parabolic rate constants decreased with increasing Al content. The scales formed on Fe-5 and –10 at.% Al were duplex, consisting of an outer layer of iron sulfide (FeS or Fe1–xS) and an inner complex scale of FeAl2S4 and FeS. Alloys having intermediate Al contents (Fe-18 and –28 at.% Al) formed scales that consisted of mostly iron sulfide and Al2O3 as well as minor a amount of FeAl2S4. The amount of Al2O3 increased with increasing Al content. The Fe 40 at.% Al formed only Al2O3 at 700°C, while most Al2O3 and some FeS were detected at T800°C. The formation of Al2O3 was responsible for the reduction of the corrosion rates.  相似文献   

13.
A series of Fe–15Cr–(2–3)Mo–(0.7–2.5)C (compositions in weight percent) steels was oxidised at 850°C and PO_2 = 5.8 × 10–20 atm, where iron oxide is unstable. All grew external Cr2O3 scales according to parabolic kinetics. Depletion of chromium from alloy subsurface regions led to dissolution of chromium-rich carbides if the original alloy carbon level was less than 1.2%. Simultaneous decarburisation caused a transformation of the original austenitic or austenoferritic structure into single-phase ferrite, stabilised by the molybdenum. Diffusion analysis of the concentration profiles within this transformed zone led to satisfactory agreement with the known diffusion coefficient for chromium in ferrite. At high carbon levels, decarburisation was slow, resulting in low chromium concentrations at the internal alloy–carbide interfaces. In these cases, the carbide dissolution did not proceed and chromia scaling rates were slowed.  相似文献   

14.
The low-temperature passivation of the oxide growth on iron is studied with the use of digital ellipsometry at a temperature of 300°C and an oxygen pressure from 10–3 to 1 mmHg. The maximum increment in the oxide thickness as a function of the oxygen pressure P a–p is observed in an hour of exposure, which indicates the active–passive transition. This passivity of iron and other metals can be caused by the multilayer and multiphase structure of the oxide film formed. As the oxygen pressure is increased, an external protective hematite layer appears on iron, and the inner quickly growing magnetite layer has no time to reach its limiting thickness. Shortening of the period of a rapid growth of magnetite, which is observed upon the increase in the oxygen pressure at the same exposure, results in the maximum of the summary thickness of the layer at a certain pressure P a–p. Hematite over the magnetite layer is usually formed as laterally spreading islets, the coalescence of which sharply decelerates the oxidation. In the time–pressure–oxide thickness plot, the areas of the low-temperature passivation of iron can be distinguished in wide ranges of temperature and pressure. The electrophysical treatment in the range of the active–passive transition sharply intensifies the oxidation of iron-based alloys and leads to the formation of layers with a substantial thickness and protective ability.  相似文献   

15.
The sulfidation kinetics and morphological development of the reaction products for a Ni-20wt.% Mo alloy exposed at 700° C to H2-H2S atm at sulfur pressures in the range 1×10–11 to 2×10–2 atm are reported. At Ps25×10–11 atm, the reaction product was Mo2S3 which grew as an external scale by parabolic kinetics. For 1×10–1024×10–10 atm, simultaneous internal precipitation and external growth of MoS2 occurred by linear kinetics. An external duplex scale was formed at sulfur pressures 2×10–822×10–2 atm in which the inner layer was a two-phase mixture of MoS2 and nickel sulfide, and the outer layer contained solid nickel sulfides and a liquid Ni-Mo sulfide phase. Catastrophic linear kinetics occurred under these latter condition.  相似文献   

16.
Oxidation, in oxygen gas at atmospheric pressure, of copper-manganese alloys (Mn content less than 40 at.%) has been investigated between 600 and 850° C. The reaction kinetics, determined by thermogravimetry, follow a parabolic law for alloys having a low manganese content (less than 10 at.% Mn) but are more complex for higher concentrations, particularly in the first stages of the oxidation process. Whereas in the early stages of oxidation the kinetics are controlled by surface reactions which accompany the formation of the different oxide layers, they are later controlled by the diffusion of a mobile species when the parabolic law is followed. In this condition an apparent activation energy may be determined from the rate constants. These energies are of the order of 120–140 kJ mol–1, comparable with that for oxidation of pure copper (134 kj mol–1), indicating a similar oxidation mechanism.The oxide layers formed were identified by cross-checking results of X-ray diffraction, electron microprobe analysis, and from glow discharge spectrometry. External layers of CuO and Cu2O formed on alloys of lower manganese concentration, evolving towards one or several mixed copper-manganese oxide layers with increasing manganese content. Under the external layers, which were weakly adherent to the sample, an internal-oxidation layer formed, which was adherent and consisted of precipitates of Mn3O4/MnO dispersed in the copper lattice. For alloys richer in manganese (36 at. % Mn) and at temperatures above 850°C (20 at.% Mn), the internal-oxidation layer evolved into two zones: MnO particles beneath a zone of Mn3U4 particles.  相似文献   

17.
Dunning  J. S.  Alman  D. E.  Rawers  J. C. 《Oxidation of Metals》2002,57(5-6):409-425
The effect of Si and Al additions on the oxidation of austenitic stainless steels with a baseline composition of Fe–16Cr–16Ni–2Mn–1Mo (wt.%) has been studied. The combined Si and Al content of the alloys did not exceed 5 wt.%. Cyclic-oxidation tests were carried out in air at 700 and 800°C for a duration of 1000 hr. For comparison, conventional 18Cr–8Ni type-304 stainless steel specimens were also tested. The results showed that at 700°C, alloys containing Al and Si, and alloys with only Si additions showed weight gains about one half that of the conventional type-304 alloy. At 800°C, alloys that contained both Al and Si additions showed weight gains approximately two times greater than the type-304 alloy. However, alloys containing only Si additions showed weight gains four times less than the 304 stainless. Further, alloys with only Si additions preoxidized at 800°C, showed zero weight gain in subsequent testing for 1000 hr at 700°C. Clearly, the oxide-scale formation and rate-controlling mechanisms in the alloys with combined Si and Al additions at 800°C were different than the alloys with Si only. ESCA, SEM, and a bromide-etching technique were used to analyze the chemistry of the oxide films and the oxide–base-metal interface, in order to study the different oxide film-formation mechanisms in these alloys.  相似文献   

18.
The oxidation behavior of 9Cr-1Mo ferrritic steels has been studied in air, oxygen, and steam at 1 atm pressure at various temperatures. Long-term experiments in air were carried out from 500–800°C by measuring the weight gains by interrupting the experiment at regular intervals of time. Short-term experiments in oxygen from 500–950°C and in air at 900 and 950°C were carried out by continuous recording of weight gain versus time in a continuousrecording thermogravimetric balance. Short-term experiments in steam were carried out using a special atmosphere furnace attached to the thermogravimetric balance. In air/oxygen, the weight gains at 700°C were lower than those at 600°C, while in steam, the weight gains at 800°C were lower than those at 700°C. This inversion phenomenon was observed for all the three steels viz. 9Cr-1Mo (high Si), 9Cr-1Mo (low Si), and 9Cr-1Mo-Nb steel. Examination of the oxide scales was carried out using SEM/EDAX, AES/ESCA, and X-ray diffraction techniques, and a mechanism is proposed for the occurrence of the inversion phenomenon.  相似文献   

19.
Ni–20 wt.% Cu was oxidized in different oxygen pressures from 1×10–5 to 1 atm at 700–1100°C. The oxidation consisted of an initial transient period in which a composite scale of NiO and Cu oxides formed, and a subsequent quasi steady-state regime during which parabolic growth of NiO determined the overall oxidation rate. Based on the oxide composition and the oxygen- pressure dependence of the parabolic rate constant, it is concluded that outward transport of Ni via vacancies determines the growth rate of the oxide during the steady-state period, either in the grain boundaries or in the lattice. The influence of Cu dissolved in NiO on the oxidation rate and the oxygen-pressure dependence of the parabolic rate constants is discussed.  相似文献   

20.
The oxidation kinetics of Rh were measured in air at 1 atm. in the temperature range 600–1000°C. The oxidation weight gain proceeds logarithmically at the lower temperatures (600°C, 650°C) followed by a transition to power law behavior at the higher temperatures (800°C). The logarithmic growth kinetics result from thickening of a hexagonal Rh2O3 scale. The transition from logarithmic to power law growth kinetics occurs in the range 700–800°C, and reflects thickening of hexagonal and orthorhombic Rh2O3 scales. Above 800°C, the growth kinetics result from thickening of a predominately orthorhombic Rh2O3 scale. At 1000°C the oxide becomes volatile, leaving the metal surface exposed.  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