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.
In coal-fired power plants using oxyfuel combustion process with carbon capture and sequestration, instead of air, a mixture of oxygen and recirculated flue gas is injected in the boiler. A series of steels were exposed to CO2-SO2-Ar-H2O gas mixtures at 600 °C for 1000 h to compare their high temperature corrosion behavior. During the corrosion process, carburization, decarburization and recrystallization were observed underneath the oxide scale depending on the gas mixture and alloy composition. The conditions that lead to carburization are not yet completely understood, but decarburization can be simulated using thermodynamic and kinetic models. In this work, the results of these simulations are compared with measured values for one of the alloys that displayed a decarburized region. Since the mobility of carbon in the scale is not known, two strategies were adopted: simulation of alloy-atmosphere contact; and estimation of the carbon flux to produce the observed decarburization. The second approach might give an insight on how permeable to carbon the scale is.  相似文献   

2.
This study focuses on the oxidation behavior of commercially available HR120 in air at 1,050 °C from 30 min to 100 h. The oxidation kinetics were first studied by thermogravimetry and isothermal exposure. The oxidation products were fully characterized using ex and in situ X-ray diffraction (XRD) and FEG-SEM observations. HR120 experienced at 1,050 °C a non protective transient stage and formed a multilayered oxide scale (SiO2–Cr2O3–XCr2O4 with X = Mn and/or Fe, Ni). A series of complementary characterization methods (gold and isotopic marker experiments, photoelectrochemistry (PEC)) were implemented to elucidate the oxidation mechanism. The study identified a n-type semi-conductivity accompanied by an inward growth of the scale. Thus, assuming that diffusion in the oxide scale controlled chromia-scale growth, the oxygen vacancy was the major point defect governing the solid state transport. This result was attributed to the presence of a MnCr2O4 spinel layer at the top of chromia that strongly decreased the oxygen pressure at the interface spinel/chromia.  相似文献   

3.
Fe–Ni–Cr–Co–P alloys were exposed to 17.3% CO2–H2 gas mixtures to investigate the oxidation of minor elements in metallic alloys in the early solar system. Reaction temperatures varied between 700 and 1000 °C. Gas-phase equilibrium was attained at 800, 900, and 1000 °C, yielding H2–H2O–CO–CO2 gas mixtures. Experiments at 700 and 750 °C did not achieve gas-phase equilibrium and were performed in H2–CO2 gas mixtures. Reaction timescales varied from 1 to 742 h. The experimental samples were characterized using optical microscopy, electron microprobe analysis, wavelength-dispersive-spectroscopy X-ray elemental mapping, and X-ray diffraction. In all experiments Cr experiences internal oxidation to produce inclusions of chromite (FeCr2O4) and eskolaite (Cr2O3) and surface layers of Cr-bearing magnetite [(Fe,Cr)3O4]. At 900 and 1000 °C, P is lost from the alloy via diffusion and sublimation from the metal surface. Analysis of P zoning profiles in the remnant metal cores allows for the determination of the P diffusion coefficient in the bulk metal, which is constant, and the internally oxidized layer, which is shown to vary linearly with distance from the metal surface. At 800 and 900 °C, P oxidizes to form a surface layer of graftonite [Fe3(PO4)2] while at 700 and 750 °C P forms inclusions of the phosphide-mineral schreibersite [(Fe,Ni)3P].  相似文献   

4.
Kai  W.  Chang  M. T.  Liu  C. D.  Lee  T. W. 《Oxidation of Metals》2002,57(5-6):523-547
The corrosion behavior of pure Y and two Y–Al alloys containing 5 and 10 wt.% Al was studied over the temperature range 800–950°C in a H2/H2S/H2O gas mixture. Both alloys had the two-phase structure of Y+Y2Al. With the exception of Y–10Al, for which a kinetics inversion was observed between 800°C and higher temperatures (T 850°C), the parabolic rate constants generally increased with increasing temperature, but decreased with increasing Al content. The scales formed on pure Y and the Y–Al alloys were single but heterophasic, consisting of mostly Y2O3 and minor Y2O2S. XRD results showed no evidence of Al2O3 and pure sulfides. The formation of Y2O3 and Y2O2S on Y–10Al at 800°C resulted in a subsurface phase transformation from Y+Y2Al to YAl2 and broke the structural integrity of the scale, being responsible for the fast corrosion rate.  相似文献   

5.
Rapid formation of an α-Al2O3 scale on Fe–50 at.%Al by pure metal thin coatings of Ni, Al, Ti, Cr or Fe was investigated, and the effects of those elements on Al2O3-scale evolution were assessed. The oxidation behavior of samples with and without coatings could be divided into two groups: the samples with/without Ni and Al, and those with Ti, Cr and Fe. The mass gains of samples coated with Al and Ni were almost the same as that of non-coated Fe–50 at.%Al alloy. The mass gains of samples coated with Ti, Cr, and Fe were much lower than that of the Fe–50 at.%Al alloy. A stable α-Al2O3 scale was found to develop from the beginning of oxidation on the samples coated with Ti, Cr and Fe. However metastable θ-Al2O3 remained after long-time oxidation of non-coated and Ni- and Al-coated samples. The direct α-Al2O3 scale formation on the samples with Cr or Fe coatings was speculated to be due to sympathetic nucleation of α-Al2O3 on the surface of Al-supersaturated Fe2O3 for Fe-coated sample, and composition changes from (Cr,Al)2O3 to (Al,Cr)2O3 for the Cr-coated sample. Initial formation of an oxide having a corundum structure was inferred to provide a nucleation site for precipitation of α-Al2O3 without prior formation of a metastable Al2O3 scale.  相似文献   

6.
A Rapidly Solidified Powder (RSP) metallurgical FeCrAl alloy, Kanthal APMT, was exposed in dry and humid O2 for 72 h at 900–1,100 °C. The formed oxide scales were characterized using gravimetry in combination with advanced analysis techniques (SEM, EDX, TEM, XRD, AES and SIMS). The oxide scales were at all exposures composed of two-layered α-Al2O3 scales exhibiting a top layer of equiaxed grains and a bottom layer containing elongated grains. A Cr-rich zone, originating in the native oxide present before exposure, separated these two layers. The top α-Al2O3 layer is suggested to have formed by transformation of outwardly grown metastable alumina, while the inward-grown bottom α-Al2O3 layer had incorporated small Zr-, Hf- and Ti-rich oxide particles present in the alloy matrix. The scale also contained larger Y-rich oxide particles. Furthermore, in the temperature range studied, the presence of water vapour accelerated alloy oxidation somewhat and affected scale morphology.  相似文献   

7.
8.
To evaluate the suitability of HR3C and 22Cr–25Ni–2.5Al AFA steels as the heat-resistant alloys, the oxidation behavior of them was investigated in air at 700, 800, 900 and 1000 °C. The evolution of oxide layer on the surface and subsurface was investigated using a combination of compositional/elemental (SEM, EDS) and structural (XRD, GDOES) techniques. A dense and continuous Cr2O3 healing layer on the HR3C was formed at the temperature of 700 or 800 °C, but the Cr2O3 oxide film on HR3C was unstable and partly converted into a less protective MnCr2O4 with the increase in temperature to 900 or 1000 °C. The composition and structure of oxide film of 22Cr–25Ni–2.5Al AFA steels are significantly different to the HR3C alloys. The outer layer oxides transformed from Cr2O3 to Al-containing oxides, leading to a better oxidation resistance at 700 or 800 °C compared to HR3C. Further, the oxide films consist of internal Al2O3 and AlN underneath the outer loose layer after 22Cr–25Ni–2.5Al AFA oxidized at 900 or 1000 °C. It can be proved that the internal oxidation and nitrogen would make 22Cr–25Ni–2.5Al AFA steels have worse oxidation resistance than HR3C alloys at 900 or 1000 °C.  相似文献   

9.
To clarify the oxidation mechanism of Cu2O to CuO, Cu2O oxidation was studied at 600–1050 °C under 1atm O 2. The Cu2O specimens were prepared through completely oxidizing 99.99999 and 99.5 pure copper at 1000°C in an Ar + 1 O 2 atmosphere. The oxidation kinetics of Cu2O specimens prepared from both purity levels followed the logarithmic law, not the parabolic law or the cubic law as reported in the literature. The activation energy for Cu2O oxidation is relatively high in the lower-temperature range, but becomes very small or even negative at higher temperatures. The logarithmic oxidation rate law can be explained by Davies et al.s model related to grain-boundary diffusion in the oxide layers. The very small or negative activation energies in the higher-temperature range can be attributed to the very small thermodynamic driving force and the fast lateral growth of CuO grains related to a sintering effect. The influence of small amount of impurities is also discussed.  相似文献   

10.
The oxidation behavior of three Fe–Si alloys containing approximately 5, 9 and 13 at.% Si has been studied at 600–800 °C under 1 atm O2. Fe–5Si and Fe–9Si followed multi-stage parabolic kinetics, while Fe–13Si showed more complex kinetics at all temperatures. The increase in Si content resulted in a transition from the internal oxidation of Si beneath an external scale of iron oxides to the exclusive external formation of silica. The critical contents required for the transitions between the various possible oxidation modes of the binary Fe–Si alloys were calculated and compared with the experimental results. A thermodynamic mechanism for the formation of Fe-rich oxide nodules observed in the oxidation of Fe–5Si and Fe–9Si has been proposed.  相似文献   

11.
The oxidation behavior of three commercial superalloys, CrMoV, F91 and Mar-M247, was studied at 550 °C in laboratory air for 1000 h. Mar-M247 superalloy showed the best oxidation resistance, which is attributed to the formation of a scale rich in Cr2O3 and Al2O3, followed by F91 and CrMoV. A thick duplex oxide formed on CrMoV alloy and spallation was observed. The results for CrMoV alloy showed that calculated Fe diffusion in magnetite was 200 times faster than literature values for Fe diffusion in Fe3O4, which is attributed to grain-boundary diffusion and the effect of impurity on diffusion. F91 initially formed a protective chromium-rich oxide layer followed by formation nodules, leading breakaway oxidation. The oxide nodules consisted of a duplex structure with different morphologies and oxide phases from duplex oxide scale in CrMoV.  相似文献   

12.
13.
The oxidation of four Ni–10Cr–ySi–4Al alloys has been studied in 1 atm O2 at 900 and 1000 °C to examine the effects of various Si additions on the behavior of the ternary alloy Ni–10Cr–4Al, which during an initial stage formed external NiO scales associated with an internal oxidation of Cr + Al, later replaced by the growth of a chromia layer at the base of the scale plus an internal oxidation of Al. The addition of 2 at.% Si was able to prevent the oxidation of nickel already from the start of the test, but was insufficient to form external alumina scales at 1000 °C, while at 900 °C alumina formed only over a fraction of the alloy surface. At 1000 °C the addition of 4 at.% Si produced external chromia scales plus a region of internal oxidation of Al and Si, a scaling mode which formed over a fraction of the alloy surface in combination with alumina scales also by oxidation at 900 °C. Conversely, the presence of about 6 at.% Si produced external alumina scales over the whole sample surface at 900 °C, but only over about 60 % of the alloy surface at 1000 °C. The changes in the oxidation modes of the ternary Ni–10Cr–4Al alloy produced by Si additions have been interpreted by extending to these quaternary alloys the mechanism of the third-element effect based on the attainment of the critical volume fraction of internal oxides needed for the transition to the external oxidation of the most-reactive-alloy component, already proposed for ternary alloys.  相似文献   

14.
A low carbon, low silicon steel was reacted with flowing N2–H2–H2O gases at temperatures of 800–1,200?°C, to produce scales of fast-growing wüstite, the only stable iron oxide under these conditions. Scaling kinetics were parabolic, after an initial period of linear reaction. The parabolic rate constants measured in the range 800–1,100?°C were two orders of magnitude lower than values predicted from Wagner’s diffusion theory, and activation energies were higher than expected. At 1,200?°C, however, the measured parabolic rate constant was in agreement with prediction. The initial period of linear kinetics was extensive at 1,100?°C, and rates are in agreement with those predicted for surface reaction rate control. Because the surface reaction is much less sensitive to temperature than the solid-state diffusion process, the extent of the linear kinetic regime is smaller at lower temperatures.  相似文献   

15.
Metal dusting of Fe–Ni alloys was investigated in a CO–H2–H2O–Ar gas corresponding to a C = 19.6 at 650 °C. Thermogravimetric analysis showed that increasing the nickel content in the alloy decreased the initial rate of carbon uptake. A uniform Fe3C scale formed on pure iron, a layer with mixed structures of Fe3C, γ and α-Fe developed on ferritic Fe–5Ni, and small amounts of Fe3C developed at the surface of an austenite layer grown on two-phase (α + γ) Fe–10Ni. At nickel levels above 10%, no carbide appeared. These observations are shown to be broadly consistent with local equilibrium according to the Fe–Ni–C phase diagram. However, the failure of higher nickel austenitic alloys to form the (Fe,Ni)3C expected at high carbon activities indicates a barrier to nucleation and growth of this phase. Graphite deposition was catalysed by (Fe,Ni)3C on ferritics and by the metal itself on austenitics. The rates of carbon deposition on Fe–60Ni corresponded to the existence of three parallel and independent paths: the synthesis gas, the Boudouard and the carbon methanation reactions.  相似文献   

16.
Kai  W.  Chang  M. T.  Bai  C. Y. 《Oxidation of Metals》2001,56(3-4):191-214
The corrosion behavior of three Ti–Al intermetallics containing 20, 30, and 40 wt.% Al was studied over the temperature range 800–1000°C in a H2/H2S/H2O gas mixture. Ti–20Al and Ti–40Al alloys had the single-phase structure of Ti3Al and TiAl, respectively, while Ti–30Al was a two-phase mixture of Ti3Al+TiAl. The corrosion kinetics followed the parabolic rate law in all cases, regardless of temperature and alloy composition. The parabolic rate constants increased with increasing temperature, but decreased with increasing Al content. The Ti–40Al alloy exhibited the best corrosion resistance among all alloys studied. The scales formed on Ti–Al intermetallics were heterophasic and duplex, consisting of an outer-scale layer of pure -TiO2 and an inner layer of -TiO2 with minor amounts of -Al2O3 and Til-xS. The amount of -Al2O3, which increased with increasing Al content, is responsible for the reduction of the corrosion rates as compared with those of pure Ti oxides.  相似文献   

17.
Four ternary Cu–Zn–Al alloys containing 5 or 10 at.% Zn and 2 or 4 at.% Al plus an alloy containing 2 at.% Al and 15 at.% Zn have been oxidized at 800 °C in 1 atm O2, and their behavior has been compared with that of the corresponding binary Cu–Zn and Cu–Al alloys. For the alloy containing 4 at.% Al, which is already able to form external alumina scales, the addition of Zn is only effective in reducing the mass gain during the fast, initial-oxidation stage. Conversely, the addition of 15 at.% Zn to Cu–2Al is able to prevent the formation of external scales containing mixtures of the Cu and Al oxides, resulting in the formation of external alumina scales after an initial stage of faster rate, producing a limited third-element effect. Finally, the addition of Al to both Cu–5Zn and Cu–10Zn is able to prevent the internal oxidation of Zn, producing a kind of reversed third-element effect. Possible mechanisms for these effects are examined on the basis of general treatments concerning the scaling behavior of ternary alloys.  相似文献   

18.
Binary Fe–(1, 2, 3)Si and Fe–(2, 4, 6)Al, and ternary Fe–(2, 3)Si–(4, 6)Al alloys (all in wt%) were oxidised in Ar–20% CO2, with and without H2O, at 800 °C. All binary alloys except Fe–6Al, in all gases, formed a thin outer layer of Fe3O4, an intermediate Fe3O4 + FeO layer, an inner FeO + Fe2SiO4 (or FeAl2O4) layer and internally precipitated SiO2 (or FeAl2O4). Ternary alloys and Fe–6Al developed a protective Al2O3 layer beneath Fe2O3 in Ar–20% CO2. Water vapour affected ternary alloy oxidation only slightly, but Fe–6Al oxidized internally in high H2O-content gas, and its scale was non-protective.  相似文献   

19.
The present study investigates oxidation at 600?°C of alloy Sanicro 25 (42Fe22Cr25NiWCuNbN) in dry and wet O2 environments. The exposure time was 1–168?h. The oxidized samples were analyzed by grazing incidence X-ray diffraction, glow discharge optical emission spectroscopy, scanning electron microscopy, transmission electron microscopy and energy dispersive X-ray spectroscopy. Alloy Sanicro 25 showed protective oxidation behaviour under the present conditions. Initially, a thin and smooth corundum-type single layer base oxide formed, featuring a Cr-rich bottom part and a Fe-rich top. With time, double-layered oxide nodules form consisting of inward- and outward-growing parts. Below the oxide scale a 100–200?nm thick oxidation-affected zone formed in the alloy, which was depleted in Cr and enriched in Ni. In this region the chromium carbides and copper-rich particles present in the bulk alloy were dissolved. In O2?+?H2O environment, chromium volatilized from the surface, causing the chromium content of the oxide to be lower than after oxidation in dry O2. However, under present experimental conditions, the Cr depletion of the scale was not enough to trigger accelerated corrosion of the alloy.  相似文献   

20.
The effects of compressive stresses on the oxide-scale morphologies formed on an Fe–20Cr alloy were investigated by comparison of the oxidation behavior in air under classical conditions, i.e., without any applied mechanical stresses and under static compressive stresses, at 900 °C. The study was carried out mainly by comparisons of oxidation kinetics gained by thermogravimetric analysis (TGA), surface morphologies of oxidized specimens observed by scanning electron microscopy (SEM), oxidized products examined by X-ray diffraction (XRD). It was found that the application of compressive stresses induced an increase in oxidation rate, but a decrease of oxide grain size. When the stresses are in the range of 5–8 MPa, both chromium- and iron-oxides formed but, at other stresses, only chromia was present. In particular, there was a maximum in oxidation rate when the applied stress was 5 MPa. The paper places emphasis on analyzing the cause of this phenomenon.  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