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 46 毫秒
1.
The influence of an oxidizing molten nitrate salt (60 pct NaNO3-40 pct KNO3) on the mechanical properties of 2.25Cr-1Mo has been examined through a series of slow strain rate tests at 450°C and 525°C. By comparing fracture strain, reduction in area, and the ultimate strength of air-exposed specimens to these same parameters for specimens tested in the binary salt mixture, the susceptibility of the alloy to environmental degradation was ascertained. Exposure to the nitrate resulted in a loss of ductility as measured by either the engineering fracture strain or reduction in area at both temperatures studied. At these temperatures, the ductility loss was most pronounced at the lowest strain rates (1×10−7 s−1). In general, for all strain rates examined, the degree of ductility loss was greater at 525°C than at 450°C. Metallographic observations revealed that severe surface oxidation occurred as the result of exposure to the molten salt. These surface scales were found to be nonadherent and easily spalled. The rapid formation of these corrosion products and the inability of the material to form a protective barrier against further oxidation is consistent with the ductility loss observed in the salt-exposed specimens.  相似文献   

2.
Metal powders always contain a surface oxide layer, which is particularly tenacious in aluminum alloys. After hot pressing, this oxide coats the particle boundaries and reduces the ductility. In this article, a study of the Al-17 wt pct Cu alloy densified from rapidly solidified powder is presented. Different thermomechanical treatments were investigated to improve the ductility of this material. Channel-die (CD) forging was performed at two temperatures (430 °C and 500 °C). Eight compression runs were applied to the samples in each CD treatment. At 430 °C, three strain values per run were investigated (35, 50 and 70 pct). A bar was also extruded with a 40:1 ratio. Because of the small size of the samples, the ductility was assessed by means of the ring expansion test and analyzed by post mortem (fracture surface and cross section) observations. No ductility was measured after CD compression at 430 °C, although it appears from the fracture surface observations that increasing the strain per run has a beneficial effect. The CD compression at 500 °C and extrusion were both successful at promoting ductility, extrusion being more effective.  相似文献   

3.
Forging limits in a discontinuously reinforced aluminum (DRA) matrix composite, 2014 Al/15 vol pct A1-2O-3, were determined by compressing samples of various cylindrical geometries under different conditions of temperature, strain rate, and lubrication and measuring the limit strains attained prior to incipient crack formation. In some cases, circumferential grids were machined on the sample surface to obtain the local fracture strain states. Crack formation was caused by the secondary tensile stresses; however, crack propagation was relatively slow and somewhat more severe at 300 °C than at 400 °C. The forging limit of the composite was found to be higher at 400 °C than at 300 °C and also higher at slower strain rates. The plane-strain forging limit of the composite at 300 °C and a strain rate of 0.5 s”1 was less than 0.05, while that of the matrix was higher than 0.5. It was found that the forging limits can be influenced by the depth of the circumferential grids and can be lower than those for the smooth surface samples. Formerly Graduate Student, the University of Michigan  相似文献   

4.
Strain controlled low cycle fatigue tests have been conducted in air to ascertain the influence of strain rate(ε = 4 × 10-6'to 4 × 10-3 s-1) and temperature(T = 750/850/950 °C) on LCF behavior of Alloy 617. A strain range of 0.6 pct and a symmetrical triangular wave form were employed for all the tests. Crack initiation and propagation modes were studied. Microstructural changes that occurred during fatigue deformation were evaluated and compared with the results obtained on isothermal aging. Deformation and damage mechanisms which influence the endurance have been identified. A reduction in fatigue life was observed with decreasing ε at 850 °C and with increasing temperature at ε = 4 × 10-5 s-1. Cyclic stress response varied as a complex function of temperature and strain rate. Fatigue deformation was found to induce cellular precipitation of carbides at 750 and 850 ‡. Dynamic strain aging characterized by serrated flow was observed at 750 °C (ε = 4 × 10-5 s-1) and in the tests at higher ε at 850 °C. Strengthening of the matrix due to dynamic strain aging of matrix dislocations by precipitation of M23C6 carbides led to fracture of grain boundary carbide films formed at 750 °C, producing brittle intergranular crack propagation. At 850 °C transgranular crack propagation was observed at the higher strain rates ε≥4× 10-4 s-1. At 850 and 950 °C even at strain rates of 4 × 10-5 s-1 or lower, life was not governed by intergranular creep rupture damage mechanisms under the symmetrical, continuous cycling conditions employed. Reduction of endurance at lower strain rates is caused by increased inelastic strain and intergranular crack initiation due to oxidation of surface connected grain boundaries. formerly Guest Scientist at the De-partment for Reactor Materials of the Nuclear Research Centre, Juelich (IRW/KFA),  相似文献   

5.
The isothermal and non-isothermal oxidation kinetics of a converter vanadium slag in the presence of calcium oxide was studied using thermal analysis. The isothermal experimental data for the whole oxidation process are described in terms of the equation [1? (1?α)2/3] = kt with Ea = 20.42 kJ mol–1 at lower temperatures of 400-500 °C, and described by [(1?α)–1/3?1]2 = kt with Ea = 227.66 kJ mol–1 at temperature higher than 500 °C. In the nonisothermal oxidation study, heating rate greatly affects the oxidation process. Using a heating rate of 3 °C min–1 results in overlapping oxidations of vanadium spinel and augite over temperature range of 608-959 °C, which is described by the 3/2 order reaction. Increasing the heating rate to 5 °C min–1 or 10 °C min–1, only oxidation of vanadium spinel takes place in temperature range of 657-914 °C and 691-954 °C respectively, both described by the third order chemical reaction. As the slag particle decreases from 250 µm to 48 µm, the kinetic equation for describing the overlapping oxidation process changes from the Anti–Zhuravlev equation with internal diffusion controlling to reaction limiting equations.  相似文献   

6.
A commercial Al-6 pct Mg-0.3 pct Sc-0.3 pct Mn alloy subjected to equal-channel angular extrusion (ECAE) at 325 °C to a total strain of about 16 resulted in an average grain size of about 1 μm. Superplastic properties and microstructural evolution of the alloy were studied in tension at strain rates ranging from 1.4 × 10−5 to 1.4 s−1 in the temperature interval 250 °C to 500 °C. It was shown that this alloy exhibited superior superplastic properties in the wide temperature range 250 °C to 500 °C at strain rates higher than 10−2 s−1. The highest elongation to failure of 2000 pct was attained at a temperature of 450 °C and an initial strain rate of 5.6 × 10−2 s−1 with the corresponding strain rate sensitivity coefficient of 0.46. An increase in temperature from 250 °C to 500 °C resulted in a shift of the optimal strain rate for superplasticity, at which highest ductility appeared, to higher strain rates. Superior superplastic properties of the commercial Al-Mg-Sc alloy are attributed to high stability of ultrafine grain structure under static annealing and superplastic deformation at T ≤ 450 °C. Two different fracture mechanisms were revealed. At temperatures higher than 300 °C or strain rates less than 10−1 s−1, failure took place in a brittle manner almost without necking, and cavitation played a major role in the failure. In contrast, at low temperatures or high strain rates, fracture occurred in a ductile manner by localized necking. The results suggest that the development of ultrafine-grained structure in the commercial Al-Mg-Sc alloy enables superplastic deformation at high strain rates and low temperatures, making the process of superplastic forming commercially attractive for the fabrication of high-volume components.  相似文献   

7.
8.
The compositional changes in Al-Li-(Mg)-(Cu) alloys induced by oxidation at high temperature (450 °C to 570 °C) were investigated by secondary ion mass spectrometry (SIMS). It was found that the alloy surface beneath the oxide layer was depleted in both Li and Mg as a consequence of the selective oxidation of these elements, whereas Cu concentration was nearly constant or slightly increased in the affected zone. The measured-concentration profiles of Li and Mg were modeled using a diffusion equation to obtain diffusion data for the alloys. The depletion profiles also provided information regarding the interfacial-alloy composition and the depletion depth. The effect of alloying elements on the oxidation and depletion behavior is discussed. Secondary ion mass spectrometry data were quantified using the relative sensitivity factor method, and the quantification procedure is described in detail.  相似文献   

9.
Subsurface void formation in oxide dispersion strengthened MA-754 caused by high temperature oxidation was investigated at temperatures of 1100, 1150, and 1200 °C for times of 1, 10, 50, and 100 hours. Material exposed at 1200 °C was examined using microprobe, SEM, and optical microscopy techniques. After exposure in air at 1200 °C for 100 hours, chromium depletion by as much as 10 wt pct was observed near the surface, and voids of various sizes up to 15 μm in diameter were found to depths of 300 μm. The fraction of voids increases with exposure time and, with the exception of anomalous values near the surface, decreases with depth. The maximum area fraction of voids observed was approximately 8 pct. Correlation of the void area fraction profile with the measured chromium depletion through a diffusion analysis shows that void formation is due to vacancy injection. Similar void formation in Ni-Cr alloys without oxide dispersions suggests that void formation is not dependent upon the presence of oxide dispersions. The diffusion coefficient for chromium in MA-754 at 1200 °C was computed from microprobe data to be 4 × 10-10 cm2 per second.  相似文献   

10.
The effect of deformation temperature in the range 250° to 500 °C on the tensile ductility of a 2014 aluminum alloy was measured in constant extension-rate tensile tests, generally with an initial strain rate of 2.6 X 10-3 s-1. Elongation at fracture increased with increasing temperature in the range 250° to 450 °C, reaching a maximum value of ~180 pct, but in stretching above ~450 °C ductility was reduced by void growth. Below ~350 °C the limiting elongations of solution-treated sheets were much inferior to those of annealed samples. Relatively poor performance in the solution-treated condition was associated with rapid hardening during the first few percent extension followed by overaging at higher strains. When solution-treated sheets were stretched above 400 °C, recovered structures and precipitation on a coarse scale developed at an early stage of straining and the stretching limits were not much inferior to those of annealed sheets at temperatures above 400 °C. The results are discussed in terms of the contributions made to necking resistance by rate-dependent and strain-dependent components of flow strength.  相似文献   

11.
The influence of grain size on the fatigue lives was investigated for eight kinds of austenitic stainless steels with the grain size numbers from 9 to 1. Fatigue tests were carried out at 600 and 700 °C under triangular wave shapes at strain rates of 6.7 × 10-3/s and 6.7 × 10-5/s, respectively, and under truncated wave shape with 30 m;n hold-time at tension side. When a strain rate was 6.7 × 10-3/s at both 600 and 700 °C, the fracture modes were always transgranular, and the fatigue lives scarcely depended on the type of steels or the grain size. When a strain rate was 6.7 × 10-5/s at 600 °C, the fracture modes changed from a dominantly transgranular mode to a completely intergranular one and the fatigue lives decreased with decreasing the grain size number. When a strain rate was 6.7 X 10-5/sVs at 700 °C, grain size dependence of the fatigue lives was divided into two groups of the steels depending on the type of steel. The fracture modes of some types of the steel were completely intergranular, and others mixed. In hold-time tests, the grain size dependence of the fatigue lives was similar to that in the tests of triangular wave shape at a strain rate of 6.7 × 10-5/s.  相似文献   

12.
The workability of 304L austenitic stainless steel has been investigated using torsion testing at temperatures from 20 °C to 1200 °C and strain rates of 0.01 and 10.0s -1 For the lower strain rate, temperature changes due to deformation heating were minimal, and failure was found to be fracture controlled at all temperatures. As for many other metals, the 304L exhibited a ductility minimum at warm-working temperatures. For the higher strain rate, failure was controlled by flow-localization processes at 20 °C and 200 °C. At these temperatures, flow softening resulting from deformation heating was deduced to be the principal cause of flow localization. A model to predict the strain at the onset of localization was developed and applied successfully to the 304L results. For high strain-rate torsion tests at 400 °C and above, failure was fracture controlled as in the low strain-rate tests, and the ductilities were shown to be correlated to those at the lower strain rate through the Zener-Hollomon parameter by employing an activation energy derived from flow-stress data.  相似文献   

13.
Total strain control fatigue tests with a 120-second hold period at either peak compressive or tensile strain were conducted on annealed 2.25 pct Cr 1 pct Mo steel. Tests were performed at the total strain range of 1.0 pct at 500 °C or 600 °C in air, 1.3 Pa (10−2 torr) or 1.3 × 10−3 Pa (10−5 torr) vacuum. The nature of the hold and the environment affect fatigue life and surface crack patterns. A compressive hold is more deleterious than a tensile hold in high-temperature air, while the reverse is true in environments in which oxidation is limited. Observations of cracks at the surface and in cross section indicate that an oxidation-fatigue interaction accounts for the damaging effect of a compressive hold in air tests. In vacuum tests, creep damage has the opportunity to accumulate and causes the tension hold to exhibit the shortest fatigue lifetime.  相似文献   

14.
Alkmin  L. B.  Chaia  N.  Utada  S.  Cormier  J.  Baldan  R.  Coelho  G.  Nunes  C. A. 《Metallurgical and Materials Transactions A》2021,52(6):2589-2600

The present investigations focused on the thermal oxidation of two variants of MAR-M246 alloy having the same contents of Ta and Nb in at. pct, considering the effects of total replacement of Ta by Nb. The alloys were produced by investment casting using high purity elements in induction furnace under vacuum atmosphere. The alloys were oxidized pseudo-isothermally at 800 °C, 900 °C and 1000 °C up to 1000 hours under lab air. Protective oxidation products growing on the surface of the oxidized samples were mainly Al2O3, Cr2O3. Other less protective oxide such as spinels (NiCr2O4 and CoCr2O4) and TiO2 were also detected as oxidation products. The conventional alloy exhibited slight internal oxidation at 800 °C and an enhanced resistance at 900 °C and 1000 °C. The Nb-modified alloy presented an exacerbated internal oxidation and nitridation at 900 °C and 1000 °C and an enhanced resistance at 800 °C. At 1000 °C, Nb-modified alloy was particularly affected by excessive spalling as the main damage mechanisms. From a kinetic point of view, both alloys exhibit the same behavior at 800 °C and 900 °C, with kp values typical of alumina forming alloys (2 × 10−14 to 3.6 × 10−13 g2 cm−4 s−1). However, Ta modified alloys exhibited superior oxidation resistance at 1000 °C when compared to the Nb modified alloy due to better adherence of the protective oxide scale.

  相似文献   

15.
Aluminum ion implantation of vanadium, niobium, and tantalum improved the metals’ oxidation resistances at 500 °C and 735 °C. Implanted vanadium oxidized only to one-third the extent of unimplanted vanadium when exposed at 500 °C to air. The oxidative weight gains of implanted niobium and tantalum proved negligible when measured at 500 °C and for times sufficient to fully convert the untreated metals to their pentoxides. At 735 °C, implantation of vanadium only slightly retarded its oxidation, while oxidative weight gains of niobium and tantalum were reduced by factors of 3 or more. Implanted niobium exhibited weight gain in direct proportion to oxidation time squared at 735 °C. Microstructural examination of the metals implanted with selected fluences of the 180 kV aluminum ions showed the following. The solubility limit of aluminum is extended by implantation, the body centered cubic (bcc) phases being retained to ~60 at. pct Al in all three metals. The highest fluence investigated, 2.4 × 1022 ions/m2, produced an ~400-nm layer of VAl3 beneath the surface of vanadium, and ~300-nm layers of an amorphous phase containing ~70 at. pct Al beneath the niobium and tantalum surfaces. All three metals, implanted to this fluence and annealed at 600 °C, contained tri-aluminides, intermetallic compounds known for their oxidation resistances. Specimens implanted to this fluence were thus selected for the oxidation measurements.  相似文献   

16.
The constitutive flow behavior of a metal matrix composite (MMC) with 2124 aluminum containing 20 vol pct silicon carbide particulates under hot-working conditions in the temperature range of 300 °C to 550 °C and strain-rate range of 0.001 to 1 s-1 has been studied using hot compression testing. Processing maps depicting the variation of the efficiency of power dissipation given by [2m/(m + 1)] (wherem is the strain-rate sensitivity of flow stress) with temperature and strain rate have been established for the MMC as well as for the matrix material. The maps have been interpreted on the basis of the Dynamic Materials Model (DMM). [3] The MMC exhibited a domain of superplasticity in the temperature range of 450 °C to 550 °C and at strain rates less than 0.1 s-1. At 500 °C and 1 s-1 strain rate, the MMC undergoes dynamic recrystallization (DRX), resulting in a reconstitution of microstructure. In comparison with the map for the matrix material, the DRX domain occurred at a strain rate higher by three orders of magnitude. At temperatures lower than 400 °C, the MMC exhibited dynamic recovery, while at 550 °C and 1 s-1, cracking occurred at the prior particle boundaries (representing surfaces of the initial powder particles). The optimum temperature and strain-rate combination for billet conditioning of the MMC is 500 °C and 1 s-1, while secondary metalworking may be done in the super- plasticity domain. The MMC undergoes microstructural instability at temperatures lower than 400 °C and strain rates higher than 0.1 s-1.  相似文献   

17.
A fracture mechanics approach was used to determine the plane strain fracture toughness (K IC) of a mechanically alloyed Al-9Ti 20 vol pct cobalt sol-gel-coated SiC particle-reinforced composite. Processing defects consisting of clumped SiC particulate, bonded by the sol-gel, initiated failure in tensile tests. The defects were measured and the fracture toughness was calculated using the Irwin relation. The value ofK IC for the as-received material was determined to be equal to 4.7 MPa·m1/2 at room temperature. Annealing the material for 120 hours and 400 hours at 500 °C increased the fracture toughness. This can be attributed to coarsening of an Al3Ti strengthening phase. Tensile tests conducted at 200 °C show thatK IC decreases at that temperature for each annealing condition. The sensitivity to the presence of the defects is greatest for samples annealed at 500 °C for 120 hours. The effect of the defects on the failure mechanism of the composite material as a function of temperature was determined. At room temperature, the Co/SiC processing defects provide low-energy paths for crack propagation; at 500 °C, the defects serve as void nucleation sites.  相似文献   

18.
The flow and fracture of MP35N (35 Co, 35 Ni, 20 Cr, 10 Mo) has been studied by uniaxial com-pression and plane strain bending in the temperature range 1000 to 1200 °C and strain rate range 0.01 to 10 s•1. This covers the normal bar rolling production conditions (∼1100 °C and 1 to 5 s“•1). The strain to fracture in plane strain bending was found to increase with increasing strain rate, roughly coinciding with the increase of the strain to the peak stress in the flow curves. Within most of the temperature and strain rate ranges investigated and under plane strain bending deformation conditions, microvoid nucleation was found to be concurrent with or greatly enhanced by the onset of dynamic recrystallization. Under these deformation conditions, flow concentration or localization along the soft layers of newly recrystallized grains oriented along the maximum shear stress directions near the surface generated microvoid nucleation and damage, in effect overriding the stress relieving and crack isolation effects normally associated with the occurrence of dynamic recrystallization. As the tem-perature was decreased toward 1000 °C and the strain rate was increased toward 10 s•1, an apparent transition to a microvoid nucleation mode by wedge cracking was observed, even at the maximum rate of 10 s•1. A further decrease in deformation temperature to 900 °C at a strain rate of 10 s•1, however, removed all evidence of microvoid nucleation (of the wedge type or otherwise) as well as any trace of dynamic recrystallization within the maximum strain imposed in the plane strain bending tests.  相似文献   

19.
The interaction between fatigue and oxidation has been studied in IN 100,* a cast nickel based su-peralloy tested in laboratory air at 1000 °C. The effect of fatigue cycling on oxidation was studied by quantitative metallography on polished specimens which were oxidized in a furnace and on strain cycled low cycle fatigue specimens. Thickness measurements have shown that matrix oxidation obeys a parabolic kinetics and is strongly enhanced by fatigue cycling. Measurements of oxide spikes have shown that interdendritic oxidation obeys at 1/4 kinetics and is weakly affected by cyclic straining. A phenomenological equation was proposed which accounts for this interaction through fatigue induced fracture events in the oxide scale. Analysis of the distribution histograms shows that the actual life to crack initiation is identical to the number of cycles to break the matrix oxide scale in the most critical areas. Formerly with the Centre des Matériaux, Ecole des Mines de Paris.  相似文献   

20.
Hot ductility and fracture mechanisms of a C-Mn-Nb-Al steel   总被引:1,自引:0,他引:1  
Hot-ductility tests of a C-Mn-Nb-Al steel were performed in a tensile machine at different strain rates of 1×10−4, 3×10−4, 1×10−3, and 3×10−3 s−1 and at temperatures of 650 °C, 710 °C, 770 °C, 840 °C, 900 °C, 960 °C, and 1020 °C, which are close to the continuous casting conditions of steel. Fracture surfaces were examined using a scanning electron microscope. It was found that low strain rates and coarse austenitic grains decrease hot ductility. At all test temperatures, when the strain rate decreases, the hot ductility also decreases because the void growth mechanism predominates over void nucleation, giving time for nucleated cracks to grow. This leads, finally, to the catastrophic failure. The minimum hot ductility was found at 900 °C for all strain rates, and the fracture was intergranular. Fractographic evidence showed that the voids formed during the deformation surrounded the austenite grains, indicating that the deformation was concentrated in ferrite bands located in the same places when the testing temperature was in the two-phase field.  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