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 206 毫秒
1.
Three γ + β NiCoCrAlY alloys (a cast alloy, a laser-surface-melted (LSM) alloy, and a coating as deposited by electron beam-physical vapor deposition (EB-PVD)) with similar average composition (Ni-20Co-19Cr-24Al-0.2Y in at. pct), but with different microstructures prior to oxidation, were oxidized for 0.5 and 1 hours at 1373 K in an Ar 20 vol pct O2 atmosphere (i.e., at a partial oxygen pressure of 20 kPa). It was found that on the alloy with β precipitates larger than 20 μm, the oxide layer was nonuniform in thickness, and had a laterally inhomogeneous composition and phase constitution. In this case, the oxide layer developed on top of the γ phase was thicker than that formed on top of the β phase and consisted of a NiCr2O4/Cr2O3 outer and an α-Al2O3 inner layer. For the thinner oxide formed on top of the β phase, the outer layer was constituted of a Cr and Co containing NiAl2O4 spinel and the inner layer also consisted of α-Al2O3. For the alloys with β precipitates smaller than 3 μm, a uniform and laterally homogeneous oxide formed, consisting of a Cr and Co containing NiAl2O4 outer layer on top of an α-Al2O3 inner layer. After oxidation, Y was distributed as numerous, small precipitates within the oxide layer for a homogeneous Y distribution prior to oxidation, or as a few, very large pegs along the γ/β phase boundaries of the alloy for an inhomogeneous Y distribution prior to oxidation. The performance of the alloys upon thermal cycling was improved for smaller β precipitates and for a more homogeneous Y distribution in the alloy prior to oxidation.  相似文献   

2.
A chemical vapor deposition (CVD) procedure was developed for preparing a high-quality α-Al2O3 coating layer on the surface of a single-crystal Ni-based superalloy using AlCl3, CO2 and H2 as precursors. A critical part of this procedure was a short-time preoxidation step (1 min) with CO2 and H2 in the CVD chamber, prior to introducing the AlCl3 precursor. Without this preoxidation step, extensive whisker formation was observed on the alloy surface. Characterization results showed that the preoxidation step resulted in the formation of a continuous oxide layer (∼50 nm) on the alloy surface. The outer part of this layer (∼20 nm) appeared to contain mixed oxides, whereas the inner part (∼30 nm) mainly consisted of α-Al2O3 grains with θ-Al2O3 as a minor phase. We observed that the nucleation of α-Al2O3 in the preoxidized layer was promoted by (1) rapid heating (10 seconds) of the alloy surface to the temperature region where α-Al2O3 was expected to nucleate; (2) the low oxygen pressure environment of the preoxidation step, which kept the rate of oxidation low; and (3) contamination of the reactor chamber with HfCl4. The preoxidized layer served as an effective diffusion barrier for mitigating the interaction with some of the alloying elements such as Co and Cr with the CVD precursors and eliminating whisker formation on the alloy surface.  相似文献   

3.
A chemical vapor deposition (CVD) procedure was developed for preparing a high-quality α-Al2O3 coating layer on the surface of a single-crystal Ni-based superalloy using AlCl3, CO2 and H2 as precursors. A critical part of this procedure was a short-time preoxidation step (1 min) with CO2 and H2 in the CVD chamber, prior to introducing the AlCl3 precursor. Without this preoxidation step, extensive whisker formation was observed on the alloy surface. Characterization results showed that the preoxidation step resulted in the formation of a continuous oxide layer (∼50 nm) on the alloy surface. The outer part of this layer (∼20 nm) appeared to contain mixed oxides, whereas the inner part (∼30 nm) mainly consisted of α-Al2O3 grains with θ-Al2O3 as a minor phase. We observed that the nucleation of α-Al2O3 in the preoxidized layer was promoted by (1) rapid heating (10 seconds) of the alloy surface to the temperature region where α-Al2O3 was expected to nucleate; (2) the low oxygen pressure environment of the preoxidation step, which kept the rate of oxidation low; and (3) contamination of the reactor chamber with HfCl4. The preoxidized layer served as an effective diffusion barrier for mitigating the interaction with some of the alloying elements such as Co and Cr with the CVD precursors and eliminating whisker formation on the alloy surface. L.M. HE, formerly Doctoral Candidate, Dept. of Chemical, Biomedical, and Material Engineering, Stevens Institute of Technology, Hoboken, NJ 07030 Y.-F. SU, formerly Doctoral Candidate  相似文献   

4.
The codeposition of alumina and titania with copper   总被引:1,自引:0,他引:1  
The codeposition of Al2O3 and TiO2 from the acid copper electrolyte was investigated with respect to solution pH, particle concentration in the electrolyte, addition agents and the nature of the crystalline phase of the oxide particles. The incorporation of oxides was found possible without the need of addition agents, provided the oxide particles are of the favorable crystalline form as α-Al2O3 and rutile TiO2. A technique is reported whereby the γ-Al2O3 particles are partially transformed to the α-form making the oxide suitable for codeposition.  相似文献   

5.
A ∼ 150-nm-thick coating layer consisting of α-Al2O3 as the major phase with a minute amount of θ-Al2O3 was deposited on the surface of a single-crystal Ni-based superalloy by chemical vapor deposition (CVD). Within 0.5 hours of oxidation at 1150 °C, the resulting thermally grown oxide (TGO) formed on the coated alloy surface underwent significant lateral grain growth. Consequently, within this time scale, the columnar nature of the TGO became established. After 50 hours, a network of ridges was clearly observed on the TGO surface instead of equiaxed grains typically observed on the uncoated alloy surface. Comparison of the TGO morphologies observed with and without the CVD-Al2O3 layer suggested that the transient oxidation of the alloy surface was considerably reduced. Also, the CVD-Al2O3 layer significantly reduced the growth rate of the TGO and improved its spallation resistance, while slowing the internal oxidation of Ta-rich areas that were present in the superalloy as-casting defects. These results demonstrated that this thin α-Al2O3 coating could be used as a means of favorably altering the TGO morphology and growth kinetics for no bond coat thermal barrier coating (TBC) applications.  相似文献   

6.
A ∼150-nm-thick coating layer consisting of α-Al2O3 as the major phase with a minute amount of Φ-Al2O3 was deposited on the surface of a single-crystal Ni-based superalloy by chemical vapor deposition (CVD). Within 0.5 hours of oxidation at 1150°C, the resulting thermally grown oxide (TGO) formed on the coated alloy surface underwent significant lateral grain growth. Consequently, within this time scale, the columnar nature of the TGO became established. After 50 hours, a network of ridges was clearly observed on the TGO surface instead of equiaxed grains typically observed on the uncoated alloy surface. Comparison of the TGO morphologies observed with and without the CVD-Al2O3 layer suggested that the transient oxidation of the alloy surface was considerably reduced. Also, the CVD-Al2O3 layer significantly reduced the growth rate of the TGO and improved its spallation resistance, while slowing the internal oxidation of Ta-rich areas that were present in the superalloy as-casting defects. These results demonstrated that this thin α-Al2O3 coating could be used as a means of favorably altering the TGO morphology and growth kinetics for no bond coat thermal barrier coating (TBC) applications. Y.-F. SU, formerly Doctoral Candidate  相似文献   

7.
Isothermal oxidation behavior of Ti-48.6 at. pct Al alloy was studied in pure dry oxygen over the temperature range 850 °C to 1000 °C. The oxidation was essentially parabolic at all temperatures with significant increase in the rate at 1000 °C. Effective activation energy of 404 kJ/mol was deduced. The oxidation products were a mixture of TiO2 (rutile) and α-Al2O3 at all temperatures. An external protective layer of alumina was not observed on this alloy at any of the temperatures studied. A layered structure of oxides was formed on the alloy at 1000 °C.  相似文献   

8.
The oxidation behavior of some Ni-Cr-Al alloys at high temperatures   总被引:1,自引:0,他引:1  
Oxidation of ternary Ni-Cr-Al alloys containing different Cr/Al ratios has been studied in the temperature range 800° to 1300°C. Most of the studies were performed in 1 atm oxygen or air, but the oxygen pressure dependence for one of the alloys was also investigated. The experimental methods included thermogravimetric measurements of oxidation rates and studies on reacted specimens by means of X-ray diffraction, metallographic techniques, electron microprobe analysis, and electron microscopy. In general, the oxidation rates decrease faster with time than that for an ideal parabolic behavior. The major reaction products were NiO, Cr2O3,α-Al2O3, and Ni(Cr,Al)2O4. The relative amounts of these were a function of composition, temperature, oxygen pressure, and reaction time. The Ni-9Cr-6Al alloy has the best oxidation resistance due to the formation ofα-Al2O3 at all temperatures investigated. The oxidation mechanism of the alloy is discussed.  相似文献   

9.
The oxidation kinetics of TiAl intermetallic at 500–900 °C in air is studied using a gravimetric method, and the phase composition of the scale is studied using an x-ray phase analysis. At t > 600 °C, the kinetics of oxidation is described by a parabolic equation. The oxides TiO2 (rutile), γ-Al2O3, α-Al2O3, Ti2O3 are found in the scale. It is shown that at the first stage the γ-Al2O3 and low-titanium oxides form on the sample surface at t < 70 °C. At t ≥ 850 °C, the Ti2O3 forms on the external surface of the scale, TiAl3 is found in the sublayer at the alloy/scale interface. It is shown that at t ≤ 800 °C the process is controlled by oxygen diffusion. At t > 800 °C, the oxidation mechanism changes: counterdiffusion of titanium ions through interstitial sites in TiO2 lattice occurs.  相似文献   

10.
The structure of transient scales formed on pure, Y-doped, and Zr-doped NiCrAI alloys was examined by transmission electron microscopy. Oxidation for 0.1 hour in 1100 °C air produced many of the features observed in mature α-Al2O3 scales, but on a much finer degree: randomly oriented 0.1 to 0.2 μm grains, dispersed porosity decreasing in size and amount toward the oxide-metal interface, and indications of strain and deformation. Other layers in the scale contained structures which preceded the random α-Al2O3 layer: γ-Al2O3, α-(Al, Cr)2O3, or Ni(Al, Cr)2O4 oxides which were composed of 0.1 μm subgrains having nearly the same crystallographic orientation. These layers were densely populated with internal precipitates and Moiré patterns. The underlying metal structures showed evidence of plastic flow (dislocations) due to growth stresses in the oxide and recovery of these interface dislocations into low energy networks. The formation of coherent layers of aluminum-depleted phases indicated the selective removal of aluminum, even for these very short times.  相似文献   

11.
12.
Systematic inoculation experiments were carried out to study the influence of various inclusions on the nucleation of the α-Al phase in Al-Si-Fe alloys at different cooling rates. The results showed that in dilute alloys, containing less than 1.5 pct Si+Fe, almost all the inclusion types have high percentages of occurrence within the α-Al phase, indicating that nucleation can be promoted on the surface of such inclusions. In a hypoeutectic Al-Si alloy containing 6.3 pct Si, the inclusion particles of MgO, TiB2, TiC, α-Al2O3, and SiC become mostly inactive nucleants and are pushed to the interdendritic regions because of the dominating poisoning effect of Si. The current results were used successfully to explain the efficiency differences between the commercial grain refiners in the hypoeutectic Al-Si alloys. Silicon is observed to preferentially segregate to the liquid-Al/inclusion interfaces so as to lower the free energy of such interfaces. A theoretical analysis of the poisoning effect of Si showed that Si segregation to the liquid/nucleant interface alters the interfacial energy balance so that the catalytic efficiency of the nucleant particles is dramatically reduced. Careful analysis showed that the poisoning effect of Si in the hypoeutectic alloy is overcome when the nucleant particles have active surface characteristics, as represented by the high catalytic potencies of γ-Al2O3, CaO, and Al4C3 particles in nucleating the α-Al phase of the hypoeutectic Al-Si alloy. Although some inclusions have comparable or higher occurrence levels than TiB2 in the α-Al phase, they cannot be used as efficient nucleants because of either their poor wettability with liquid aluminum or their chemical reactivity, which can change the alloy chemistry.  相似文献   

13.
The formation and coarsening of Al2O3 dispersoids have been investigated at 500 °C, 550 °C, and 600 °C in a mechanically alloyed (MA) extrusion of composition Al-0.35wt pct Li-1wt pct Mg-0.25wt pct C-10vol pct TiO2 for times up to 1500 hours. In the as-extruded condition, the dispersed phases included Al3Ti, Al4C3, MgO, cubic TiO (C-TiO), monoclinic TiO (M-TiO), TiO2, and a small amount of Al2O3. However, numerous Al2O3 dispersoids (various polymorphs: η, γ, α, and δ) with “block-shaped” morphology were formed after heat treatment due to reduction of C-TiO, M-TiO, and TiO2. Transmission electron microscopy (TEM) and X-ray diffraction (XRD) showed conclusively the transformation of these phases to additional Al2O3 and Al3Ti. High resolution TEM showed that the α-Al2O3 dispersoids exhibited some lattice matching with the α-Al matrix. Coalescence of the block-shaped Al2O3 dispersoids occurred after heat treatment, and Al4C3 also became attached to them. The length and width of the block-shaped Al2O3 dispersoids increased by a factor of ∼1.55 between 340 and 1500 hours at 600 °C.  相似文献   

14.
The oxygen content of liquid Ni-Mn alloy equilibrated with spinel solid solution, (Ni,Mn)O. (1 +x)A12O3, and α-Al2O3 has been measured by suction sampling and inert gas fusion analysis. The corresponding oxygen potential of the three-phase system has been determined with a solid state cell incorporating (Y2O3)ThO2 as the solid electrolyte and Cr + Cr2O3 as the reference electrode. The equilibrium composition of the spinel phase formed at the interface of the alloy and alumina crucible was obtained using EPMA. The experimental data are compared with a thermodynamic model based on the free energies of formation of end-member spinels, free energy of solution of oxygen in liquid nickel, interaction parameters, and the activities in liquid Ni-Mn alloy and spinel solid solution. Mixing properties of the spinel solid solution are derived from a cation distribution model. The computational results agree with the experimental data on oxygen concentration, potential, and composition of the spinel phase.  相似文献   

15.
The present study compares the performance of alumina coatings prepared by two different methods (micro arc oxidation (MAO) and detonation gun (D-gun) spray) on AA 6063 (Al alloy) fatigue test samples under plain fatigue and fretting fatigue loading. While MAO coating had comparable proportions of γ-Al2O3 and α-Al2O3, D-gun sprayed coating contained γ-Al2O3 with minimal quantities of α-Al2O3. MAO coating was relatively harder than D-gun sprayed coating. As both types of coated samples were ground, they exhibited almost the same surface roughness. D-gun sprayed alumina coated samples exhibited slightly higher magnitude of surface residual compressive stress compared with MAO coated specimens. Both types of coated samples experienced almost the same friction force. D-gun spray coated samples exhibited superior plain fatigue and fretting fatigue lives compared with MAO coated specimens. This may be attributed to layered structure of the D-gun sprayed coating.  相似文献   

16.
The results of electron diffraction study of the oxidation of a eutectic Pb-Bi alloy during heating at various partial oxygen pressures in the gas phase are presented. It is revealed that only the oxide phases of lead form at the initial stages of oxidation, which occurs from α-PbO2 through intermediate oxides nPbO2 · mPbO and Pb3O4 to the β-PbO modification.  相似文献   

17.
Friction and wear are studied for materials of the system TiN — AlN preliminary oxidized at 800–1100°C. It was established that thin oxide films containing Al2TiO5 and α-Al2O3, that promote a decrease in frictional wear, form on the surface of composite materials of the system TiN — AlN. Our assumptions are confirmed that the improvement in tribological properties of TiN — AlN composites is caused by forming oxide screening layers that prevent direct contact between the ceramics and steel counter-body. At high rates (V=16 m/sec) and pressure (P=2.0 MPa) the oxide films form more rapidly. Translated from Poroshkovaya Metallurgiya, Nos. 1–2(411), pp. 121–124, January–February, 2000.  相似文献   

18.
Dry milling of gibbsite has been carried out for 5 h in planetary ball mill to study the effect of mechanical activation on α-Al2O3 formation. Gibbsite undergoes phase transformation during milling and has resulted nanocrystalline boehmite after 5 h of milling. The average crystallite size and the BET surface area of the nanocrystalline boehmite resulted by 5 h milling of gibbsite are 8 nm and 140 m2/g, respectively. The nanocrystalline boehmite has shown reduction in the α-Al2O3 formation temperature as well as in the activation energy of α-Al2O3 formation. The average crystallite size of nanocrystalline boehmite derived α-Al2O3 is measured to be 100 nm by TEM analysis and the BET surface area of resulted nanocrystalline α-Al2O3 is 12 m2/g.  相似文献   

19.
The possibility of the formation of bonding between the two layers of a double-oxide film defect when held in a commercial purity liquid Al alloy was investigated. The defect was modeled experimentally by maintaining two aluminum oxide layers in contact with one another in a commercial purity Al melt at 1023 K (750 °C) for times ranging from 7 minutes to 48 hours. Any changes in the composition and morphology of these layers were studied by scanning electron microscopy (SEM) and energy-dispersive X-ray spectroscopy (EDX). The results showed that the oxide layers started to bond to one another after approximately 5 hours, and the extent of the bonding increased gradually by the holding time. The bonding is suggested to form because of the transformation of γ- to α-Al2O3. A complete bonding formed between the layers only when the oxygen and nitrogen trapped between the two layers were consumed, after approximately 13 hours. The results also confirmed that the nitrogen within the atmosphere of an oxide film defect reacts with the surrounding Al melt to form AlN at the interface of the defect and the melt.  相似文献   

20.
Interfaces of TiB2−NiAl and α-Al2O3−NiAl in TiB2/NiAl composites have been investigated by analytical electron microscopy. Although no consistent crystallographic orientation relationships have been found between NiAl and TiB2 or Al2O3, semicoherent interfaces between α-Al2O3 and NiAl have been observed by high-resolution electron microscopy (HREM) in areas where the low indexed crystallographic planes of α-Al2O3 aligned with that of NiAl. No semicoherent interfaces between NiAl and TiB2 have been observed. Silicon segregation was consistently detected by X-ray energy-dispersive spectroscopy (EDS) at the TiB2/NiAl interface region. Segregation has not been detected in the α-Al2O3−NiAl interface region. The segregation layer observed at the TiB2−NiAl interface is too thin to absorb any of the thermal residual stress.  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