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 kinetics of conventional Fe–20Cr–5Al (in mass %) foil, Al-deposited foil and Al-deposited and preoxidized foil was studied at 1373 K in air. All the foils were 50-m thick and contained minor additions of rare-earth elements. The oxide scales were observed with SEM and TEM combined with EDS and were characterized with X-ray diffractometry and electron diffraction. The deposition of Al onto the foil from the vapor phase improves oxidation resistance. The details regarding this matter were reported elsewhere. The combination of the Al deposition and the subsequent preoxidation at 1173 K for 90 ks in air further increases the oxidation resistance, i.e., the smallest parabolic rate constant among the three kinds of foils, and excellent scale adherence. Preoxidation enhances the growth of -Al2O3, which transforms to -Al2O3 during subsequent oxidation. However, such -Al2O3 grains are much larger than those formed on the conventional foil of similar chemical composition. Small closed voids and small spinel-type, oxide particles appear in -Al2O3 grains with the progress of oxidation. The former is explained in terms of the volume decrease accompanying the phase transformation and the latter by the low solubility of Fe in -Al2O3.  相似文献   

2.
Hiramatsu  N.  Stott  F. H. 《Oxidation of Metals》2000,53(5-6):561-576
Thin foils of Fe–20Cr–5Al alloys are susceptible to breakawayoxidation once the aluminum content of the substrate has fallen below somecritical value. The combined addition of 0.1 wt.% lanthanum and 0, 1, or 2wt% molybdenum has a beneficial effect on the high-temperature oxidation ofsuch foils. Lanthanum has the well-known reactive-element effect on adhesionof the protective alumina scale, thereby increasing the time to onset ofbreakaway oxidation, while, for alloys containing molybdenum, breakawayoxide spreads relatively slowly over the specimen in comparison to alloysthat contain no molybdenum. In particular, molybdenum-containing alloys areable to develop a protective Cr2O3 layer at the breakawayoxide–substrate interface. Conversely, molybdenum-free alloys form aninternal-oxide zone in the substrate adjacent to this interface, rather thana Cr2O3 layer, so breakaway oxide spreads rapidly. A martensitic phase isobserved in the substrate adjacent to the breakaway oxide formed on Fe–20Cr–5Al–La specimens, which means that the-phase has transferred to the -phase at the temperature ofthe oxidation test (1150°C). Conversely, -phase is retained inthe molybdenum-containing alloy, even after breakaway takes place, sincemolybdenum, which is a strong ferrite former, is enriched in the alloyadjacent to areas of breakaway oxide. The diffusion rate of chromium isslower in the than in the -phase so a continuouschromium-rich oxide layer, which is effective in inhibiting breakawayoxide from spreading, cannot be established at the breakawayoxide–substrate interface for the molybdenum-free alloys.  相似文献   

3.
Cueff  R.  Buscail  H.  Caudron  E.  Issartel  C.  Riffard  F. 《Oxidation of Metals》2002,58(5-6):439-455
The oxidation behavior of three commercial Fe–Cr–Al alloys, Kanthal APM, Kanthal A1, and Kanthal AF (containing alloying additions of yttrium), has been investigated during isothermal exposures in air at 1173 K. After an initial transient stage, a diffusional process appears to predominantly control the oxidation kinetics of both alloys. During the transient stage, relatively important mass gains have been registered and the presence of yttrium does not seem to have a significant effect on the oxidation rate. On the contrary, the reactive element markedly influences the parabolic oxidation rate and the composition of the oxide scale. In situ X-ray diffraction (XRD) shows that yttrium promotes the transformation of transition alumina into -Al2O3, leading to the formation of a more protective oxide scale.  相似文献   

4.
Chen  Guofeng  Lou  Hanyi 《Oxidation of Metals》2000,54(1-2):155-162
Magnetron-sputter deposition was used to produce a Ni–8Cr–3.5Al(wt.%) nanocrystalline coating on substrates of the same alloy. Theoxidation behavior of the cast Ni–8Cr–3.5Al alloy and itssputtered coating were investigated at 1000°C in air. Complex,layered-oxide scales composed of Cr2O3 outer layer,mixed spinel NiAl2O4 and NiCr2O4middle layer, and -Al2O3 inner layer were formedon the Ni–8Cr–3.5Al nanocrystalline coating during 200-hroxidation, whereas Cr2O3, with some NiCr2O4external layer with internal Al2O3, formed on the castalloy. Because of the formation of this -Al2O3inner layer on the coating, the sputtered Ni–8Cr–3.5Al coatingshowed better oxidation resistance than the cast alloy. The effect ofnanocrystallization on oxide formation is discussed. It was indicated thatthe formation of this -Al2O3 inner layer was closely related to therapid diffusion of Al through grain boundaries in the nanocrystallinecoating and the relatively high Cr content in Ni–8Cr–3.5Al.  相似文献   

5.
Kuiry  S. C.  Seal  S.  Fei  W.  Quick  N. 《Oxidation of Metals》2003,59(5-6):543-557
Fiber media composed of Fe–Cr–Al–Y alloy are being used increasingly as materials for high-temperature applications for their excellent oxidation resistance. The oxidation kinetics of Fe–Cr–Al–Y alloy fiber medium as a heat-resistant material for high-temperature applications was studied in dry air at 1073, 1188, 1255, and 1318 K. The oxidation process followed the parabolic kinetic law. The alumina-scale growth was found to be influenced by short-circuit diffusion and the presence of stresses related to oxide-scale growth. The surface of the oxide scale formed on the fiber medium was analyzed using X-ray photoelectron spectroscopy, which revealed that the outer surface of the oxide scale formed on the fiber medium composed of 12-m diameter Fe–Cr–Al–Y alloy fibers, consisted of -Al2O3, -Al2O3, and Cr-oxide. The metastable -Al2O3 subsequently partially transformed into the more stable -phase following a time-temperature-transformation relationship. The surface morphology and the cross section of the oxide scale formed on the fiber medium in the temperature range 1073–1318 K in dry air, have been studied by scanning-electron spectroscopy (SEM) and focused-ion beam, respectively.  相似文献   

6.
The oxidation behavior of Fe–20Cr–5Al–(0.5–5)Si and Fe–(12–20)Cr–(5–7)Al–(1–2)Si alloys extruded into honeycomb structures has been investigated at 1150°C in air for up to 500 hr. The oxidation weight gains decrease with increasing Si and Cr contents in the 5-Al alloys. Si additions are more efficient than Cr additions to reduce the weight gain. Increasing Si content in the 5-Al alloys suppresses the formation of an iron-chromium complex oxide, forming mullite and vitreous silica in the scale, although the location is not clearly indicated. The 5-Si alloy shows anisotropy in elongation of the honeycomb specimen during oxidation in the Fe–20Cr–5Al–xSi alloys, whereas alloying with Si and Cr does not improve the oxidation resistance of the 7-Al alloys significantly. These results are explained by Wagner's theory of a secondary getter. However, we point out additionally that the difference between Si and Cr in the Pilling-Bedworth ratio and the solubility of their oxides in the Al2O3 scale may contribute to the significant effect of Si additions. Finally, this paper demonstrates that the selected Fe–Cr–Al–Si honeycombs having walls 200 m thick show excellent oxidation resistance over 500 hr at 1150°C in air. The time to catastrophic oxidation is roughly proportional to the wall thickness in extruded honeycombs.  相似文献   

7.
The effects of hydrogen and Y_2O_3 on high-temperature oxidation of Cr in 20 mbar O2 have been studied at 900°C. Oxidation- and O2-dissociation rates were determined from gas-phase measurements. Hydrogen in Cr leads to breakdown of the oxide scale. The oxide scale on Cr–1%Y2O3 charged with hydrogen for 4 hr (resulting in Cr–1%Y2O3 with approximately 10 ppm hydrogen) is adherent to the metal substrate. The oxidation rate is similar for Cr with 1 ppm hydrogen and Cr–1%Y2O3 with 1 ppm hydrogen, but significantly lower for 4-hr H-charged Cr–1%Y2O3. The oxidation rate of Cr–5%Fe–1%Y2O3–25 ppm H is also lower than the oxidation rate of Cr–5%Fe–1%Y2O3–1 ppm H. This indicates that unless hydrogen is present, there are virtually no effects of the addition of 1% Y2O3 to Cr. Using labeled oxygen, 16,16O2 and 18,18O2, was found that at 900°C the dissociation rate of O2 is higher on Y2O3 than on Cr2O3. It is suggested that the well-known improvement of the oxidation resistance of Cr as an effect of additions of Y or Y2O3 is related to an increase in the dissociation rate of O2 and that there is a synergetic effect in combining Y2O3 and hydrogen.  相似文献   

8.
Liu  Zhenyu  Gao  Wei  He  Yedong 《Oxidation of Metals》2000,53(3-4):341-350
Studies using advanced analytical techniques indicated that the reactiveelements (RE) segregate along the oxide grain boundaries and at theoxide–alloy interface during oxidation of -Al2O3forming alloys. The segregation results in inward oxygen diffusion along theoxide grain boundaries as the predominant transport process in the oxidegrowth. The present work establishes a mathematical model based on themechanisms of inward oxygen diffusion along the grain boundaries and oxidegrain coarsening. This model has been used to describe the oxidationkinetics of Y-doped Fe–Cr–Al alloys. The results showed a muchbetter agreement with the experimental data than the parabolic rate law. Byusing this model, the exponential number for the grain coarsening of aluminascales during oxidation was calculated to be 3. The activation energyfor oxygen diffusing along the grain boundaries was 450 kJ/mol. They arealso in good agreement with values reported in the literatures.  相似文献   

9.
The alumina scales on a variety of high-temperature alloys are found to fluoresce when illuminated with light having a frequency greater than 18,000 cm–1. The fluorescence exhibits two narrow lines characteristic of chromium-doped alpha-aluminum oxide. The frequency shift of the two lines from the room-temperature, stress-free values of 14,402 cm–1 (1.786 eV) and 14432 cm–1 (1.789 eV) provides a noncontact measure of the stress in the alumina scales using the piezospectroscopic effect. In addition, the broadening of the lines is a measure of the stress gradient in the scale. The physical basis for the fluorescence technique is described together with its implementation for highspatial-resolution (2 m) measurements. As illustration, room-temperature measurements of the residual stress in scales formed at 1100°C on single-crystal NiAl, polycrystalline Ni3Al, two Fe–Ni–Cr–Al alloys, and two Ni–Al base superalloys are presented.  相似文献   

10.
Chen  Guofeng  Lou  Hanyi 《Oxidation of Metals》2000,53(5-6):467-479
A NiO-forming Ni–5Cr–5Al (at.%) alloy has been developed anddeposited as a sputtered nanocrystalline coating. The oxide formation andoxidation behavior of this coating have been studied at 1000°C inair. The oxidation rate markedly decreased with time and the oxidationkinetics obeyed the fourth power law. Complex oxide scales, consisting ofNiO, NiAl2O4 and -Al2O3,were formed during 200 hr oxidation. The outer oxide layer consisted of NiOand NiAl2O4 and an inner oxide layer of-Al2O3. The sputtered Ni–5Cr–5Alnanocrystalline coating showed good oxidation resistance due to theformation of an -Al2O3 inner layer andexcellent adhesion of the complex oxide scales.  相似文献   

11.
Phase diagrams Fe–Si–O and Cr–Si–O and potential–pH diagrams for the systems Fe–Si–H2O, Cr–Si–H2O, and alloy Fe + 25% Cr + 3% Si (-phase)–H2O at 25°C are constructed. Thermodynamics of the silicon effect on the chemical and electrochemical stability of iron–chromium alloys is discussed.  相似文献   

12.
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   

13.
Ul-Hamid  A. 《Oxidation of Metals》2002,58(1-2):41-56
A study was conducted to investigate the effect of Y addition on the isothermal-oxidation behaviors of Ni–10Cr, Ni–10Cr–0.5Y, Ni–10Cr–5Al, and Ni–10Cr–5Al–0.5Y alloys. The alloys were oxidized in air for 50 hr at 1000°C. The oxides formed on the alloys were characterized using primarily cross-sectional transmission-electron microscopy techniques along with light microscopy, scanning-electron microscopy, and X-ray diffraction. Although the Al-containing alloys showed comparatively better oxidation behavior, all alloys exhibited nonprotective scaling, as suggested by the thick oxides formed. The major component of the outer oxide was NiO. However, modified Y-containing alloys formed protective layers (i.e., -Cr2O3 for NiCrY and -Al2O3 for NiCrAlY) at the scale–alloy interface following the nonprotective scaling. The spalling resistance of the modified Y-containing alloys was better than their counterpart unmodified Y-free alloys, while their overall oxidation mechanism remained unchanged after Y addition.  相似文献   

14.
Calculations of parabolic reaction rate constants   总被引:3,自引:0,他引:3  
The oxidation kinetics of only a very limited number of pure metals or binary alloys can be described by the simplest parabolic law, m2=kpt, Thus for a transient period of faster kinetics, the steady state parabolic law is given by (m–mi)2 = kp(t–ti) when the initial weight gain mi does not contribute to steady state rate control. In such a case, a plot of the kinetics data as m vs t1/2 is inherently superior to the m2 vs t plot for an accurate determination of the steady state parabolic rate constant, as well as for the analysis of the transient, faster kinetics.  相似文献   

15.
Oxidation tests of rare-earth-modified and Ti-modified Fe–20Cr alloy foils, which are under consideration for catalytic converter supports, were performed in a synthetic exhaust-gas atmosphere (N2+H2O+CO2) between 900°C and 650°C. Between 900°C and 750°C, the rare earths had no effect on oxide growth rates while Ti increased growth rates. Oxide growth rates for the rareearth alloys at 800°C and 750°C are much lower than those found in the literature for oxidation of Fe–Cr alloys or pure Cr in O2-rich atmospheres. The slow growth rates for the rare-earth alloys agree with literature data for oxidation of stainless steels containing >20% Cr in wet atmospheres and are caused by growth of an oxide scale only one grain thick. At temperatures 700°C, Fe–20Cr alloys grow massive Fe oxides; however, this can be suppressed by adding rare earths or Ti. To ensure good oxide adherence, free sulfur must be eliminated in the alloy by tying it up with a reactive-element addition. Both Ti and the rare earths can be used to tie up S, but the rare earths are more effective. For converter applications, the optimum alloy composition may contain rare earths for good oxide adherence and a small amount of Ti to suppress growth of Fe-rich oxides.  相似文献   

16.
Several features of the microstructure and the adherence of alumina scales formed on Fe–Cr–Al and Fe–Cr–Al–Y single- and polycrystalline alloys after oxidation at 1000°C were examined. The convolutions of the scale and especially of the scale/alloy interface are thought to be the major reason of poor spallation resistance of scales on the yttrium-free alloy. The flat oxide scale on the even interface of the yttrium-doped alloy, on the contrary, exhibits excellent adherence upon cooling. Interfacial cavities observed on the Fe–Cr–Al alloy result from the scale undulation under compressive growth stresses. The shape and the number of cavities depend on the initial surface orientation and most probably reflect a balance between the interfacial energy of convoluted substrate in contact with the oxide layer and the energy of separated surfaces.  相似文献   

17.
Maréchal  L.  Lesage  B.  Huntz  A. M.  Molins  R. 《Oxidation of Metals》2003,60(1-2):1-28
The oxidation kinetics of two ODS Fe–Cr–Al alloys, PM 2000 and MA 956, were studied in oxygen and in air under isothermal conditions from 1000 to 1300°C. They both form an -alumina scale and have good oxidation resistance, without any mass loss. Although the aluminum content in these alloys is higher than the minimum Al content necessary to ensure the growth of a continuous alumina scale, an aluminum depletion occurred in the substrate. This depletion allows the determination of aluminum diffusion coefficients in the ODS alloy. This method is very original and interesting as no Al-stable isotope is available. Moreover, the evolution of the aluminum concentration in the substrate allows one to determine the lifetime of these alloys: indeed, when the aluminum content decreases and becomes lower than a critical value, alumina can no longer form, and less-stable oxides grow very rapidly compared to alumina.  相似文献   

18.
Sadique  S. E.  Mollah  A. H.  Islam  M. S.  Ali  M. M.  Megat  M. H. H.  Basri  S. 《Oxidation of Metals》2000,54(5-6):385-400
The detailed oxidation behavior of Fe–10Cr alloys, containing aluminumin the range of 2–8% by weight, was studied in pure oxygen at 1 atmpressure. The investigations were performed over the temperature range950–1050°C under cyclic conditions (3-hr cycles) in each case. Thecyclic-oxidation resistance, as measured by the specific weight-gain values,was observed to progressively improve with increasing aluminum content inthe alloy. For a particular aluminum content, however, the oxidationresistance decreased with increasing temperature. Following the initialtransient-oxidation period, a healing layer of chromia was established onthe four alloys. The lower-aluminum alloys (2–4% Al) were observedto end up with Fe-rich oxide scales under the experimental conditions atall temperatures, whereas those containing aluminum in the range of6–8% formed -Al2O3 scales.  相似文献   

19.
Hou  P.Y.  Moskito  J. 《Oxidation of Metals》2003,59(5-6):559-574
Using a 30-nm field-emission Auger spectroscopy probe, the segregation of sulfur to a growing oxide–metal interface was studied. The interfaces were formed by the oxidation of a Fe–40at.% Al alloy at 1000°C for various times. Both the oxide and the alloy sides of the interface were examined after spalling the surface Al2O3 layer in ultra-high vacuum. Results were compared with similar studies performed using conventional AES and related to scale development and the interface microstructure. Sulfur started to segregate to the interface only after a complete layer of -Al2O3 developed there, its concentration then increased slowly with further oxidation until reaching a level close to half a monolayer. Higher amounts were observed on interfacial-void surfaces, where Al and S cosegregated. The study showed that sulfur segregation to oxide–alloy interfaces depended on the type of interface, indicating possible relationships between segregation energies and interface microstructure.  相似文献   

20.
In this study, the oxidation properties of Ti3Al+Nb bulk alloys, as well as IMI 829 alloy, coated with a Ti3Al+Nb layer, have been considered. Model alloys have been prepared, with 5–25 at.% niobium contents; 50-m-thick Ti3Al+10 at.% Nb coatings have also been deposited on IMI 829 by triode sputtering. Bulk alloys and coated substrates have been exposed to cyclic and isothermal oxidation in air between 700 and 800°C. Niobium additions generally caused the oxidation rate of Ti3Al to decrease significantly. In all cases rutile is the main oxide formed. It is believed that the ability of niobium to dissolve in the rutile lattice, and therefore to lower the oxygen diffusion rate through the oxide layer, is a contribution to the observed oxidation resistance enhancement. The formation of niobium oxide has also been envisaged for this matter.  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