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 240 毫秒
1.
The development of microstructure and strength during forging in a single-phase austenitic stainless steel, 304L, was investigated by means of forward extrusion of cylindrical specimens. The temperature, strain, and strain rate of deformation were varied. A low strain rate was imparted by press forging (PF), and a high strain rate by high-energy-rate forging (HERF). Low forging temperatures produced dynamically recovered microstructures and monotonic increases in strength with increasing strain for low and high strain rates. At higher forging temperatures, the high-energy-rate-forged material exhibited softening, after the application of a critical amount of strain, as a result of static recrystallization which occurred within a few seconds after cessation of deformation. Analysis of isothermal compression test data, specifically the strain-to-peak stress associated with the onset of dynamic recrystallization, confirmed that dynamic recrystallization would not be expected for the deformation conditions imposed during forward extrusion in this study. Recrystallized grain size was found to vary uniquely with strain, initial grain size, and the Zener-Hollomon parameter. Recrystallization was much less prevalent in press-forged material and may have been affected by die chilling as well as the predominance of dynamic recovery. The variation of strength, recrystallized grain size, and extent of recrystallization with the deformation parameters, temperature and strain, are presented as a set of processing-property maps for each forging technique (έ). The findings are discussed in the context of developing process design criteria for forging alloy 304L.  相似文献   

2.
A model for microstructure evolution in adiabatic shear bands   总被引:2,自引:0,他引:2  
A mechanical subgrain rotation model is proposed to account for the recrystallized grains which have been observed to form in adiabatic shear bands in a number of materials. The model is based on a “bicrystal” approach using crystal plasticity theory to predict the evolution of subgrain misorientations. These mechanically induced rotations are shown to occur at the high strain rate associated with adiabatic shear band formation. Recrystallized grain formation is proposed to occur by the formation and mechanical rotation of subgrains during deformation, coupled with boundary refinement via diffusion during shear band cooling. This model is referred to as progressive subgrain misorientation recrystallization and appears to account for shear band microstructures in a variety of metals.  相似文献   

3.
Superalloys are being employed in more extreme conditions requiring higher strength, which requires producers to forge products to finer grain sizes with less grain size variability. To assess grain size, crystallographic texture, and substructure as a function of forging conditions, frictionless uniaxial compression testing characteristic of hot working was performed on INCOLOY 945 (Special Metals Corporation, Huntington, WV), which is a newly developed hybrid of alloys 718 and 925, over a range of temperatures and strain rates. The microstructure and texture were investigated comprehensively using light optical microscopy, electron backscatter diffraction (EBSD), electron channeling contrast imaging (ECCI), and transmission electron microscopy (TEM) to provide detailed insight into microstructure evolution mechanisms. Dynamic recrystallization, nucleated by grain/twin boundary bulging with occasional subgrain rotation, was found to be a dominant mechanism for grain refinement in INCOLOY 945. At higher strain rates, static recrystallization occurred by grain boundary migration. During deformation, duplex slip along {111} planes occurred until a stable 〈110〉 fiber compression texture was established. Recrystallization textures were mostly random but shifted toward the compression texture with subsequent deformation. An exception occurred at 1423 K (1150 °C) and 0.001 seconds−1, the condition with the largest fraction of recrystallized grains, where a 〈100〉 fiber texture developed, which may be indicative of preferential growth of specific grain orientations.  相似文献   

4.
The hot deformation behavior of nickel-base superalloy UDIMET 720 in solution-treated conditions, simulating the forging process of the alloy, was studied using hot compression experiments. Specimens were deformed in the temperature range of 1000 °C to 1175 °C with strain rates of 10−3 to 1 s−1 and total strain of 0.8. Below 1100 °C, all specimens showed flow localization as shear band through the diagonal direction, with more severity at higher strain rates. A uniform deformation was observed when testing between 1100 °C and 1150 °C with dynamic recrystallization as the major flow softening mechanism above 1125 °C. Deformation above γ′ solvus temperature was accompanied with grain boundary separation. The hot working window was determined to be in the interval 1100 °C to 1150 °C. Thermomechanical behavior of the material was modeled using the power-law, the Sellars-Tegart, and an empirical equation. The flow stress values showed a nonlinear dependence of strain rate sensitivity to strain rate. The analysis indicated that the empirical method provides a better constitutive equation for process modeling of this alloy. The apparent activation energy for deformation was calculated and its variations with strain rate and temperature are discussed.  相似文献   

5.
The deformation and dynamic recrystallization behavior of Waspaloy-ingot material with coarse, columnar grains was established using isothermal uniaxial and double-cone compression tests. Testing was conducted along different test directions relative to the columnar-grain microstructure at supersolvus temperatures (1066 °C and 1177 °C) and strain rates (0.005 and 0.1 s−1), which bracket typical ingot-breakdown conditions for the material. The flow behavior of axial samples (i.e., those compressed parallel to the columnar-grain direction) showed an initial strain-hardening transient followed by steady-state flow. In contrast, the stress-strain curves of samples upset transverse to the columnar grains exhibited a peak stress at low strains, whose magnitude was greater than the steady-state flow stress of the axial samples, followed by flow softening. The two distinct flow behaviors were explained on the basis of the solidification texture associated with the starting ingot structure, differences in the kinetics of dynamic recrystallization revealed in the double-cone tests, and the evolution of deformation and recrystallization textures during hot working. Dynamic recrystallization kinetics were measurably faster for the transverse samples as well as specimens oriented at ∼45 deg to the forging direction, an effect partially rationalized based on the initial texture and its effect on the input rate of deformation work driving recrystallization. Despite these differences, the overall strains required for dynamic recrystallization were comparable to those measured previously for fine-grain (wrought) Waspaloy. However, the Avrami exponents (∼2 to 3) were somewhat higher than those for wrought material (∼1 to 2), an effect attributable to the particle-stimulated nucleation in the ingot material.  相似文献   

6.
Dynamic variations in grain structures and grain boundary characteristics of NiCrFe-based alloy 718 during hot uniaxial compression as well as stress relaxation after the compression were investigated in this article. An electron backscatter diffraction (EBSD) technique was used for the specimens that were compressed at temperatures of 1010 °C and 1066 °C and strain rates of 0.5 and 0.005 s−1, up to a strain of 0.7. Stress relaxation was observed by keeping the upper die in position at the test temperatures as soon as the compression was completed. The variations in the CSL boundary distribution and in the misorientation angle distribution during compression and stress relaxation were thoroughly analyzed to characterize the dynamically recrystallized grain (DRX) boundaries. During deformation at a high strain rate of 0.5 s−1, dynamically recrystallized grains were formed by progressive subgrain rotation. Active dynamic recovery (DRV) at 1066 °C was inferred from the similar degree of strain softening in spite of the different fraction of dynamic recrystallization, which is supported by the high frequency of low misorientation angle boundaries. Stress relaxation was caused by a coalescence of subgrains having very small misorientation angles. Directional grain growth and a redistribution of the grain boundary character caused by the grain rotation occur during the stress relaxation, resulting in reduced total boundary energy. This article is based on a presentation made in the symposium entitled “Processing and Properties of Structural Materials,” which occurred during the Fall TMS meeting in Chicago, Illinois, November 9–12, 2003, under the auspices of the Structural Materials Committee.  相似文献   

7.
A study of the plastic deformation of several fcc metals and alloys at large strains was conducted. The purpose of this study was to take a critical look at the assumption of steady-state flow at low and intermediate temperatures. For this purpose, large strain data were obtainedvia torsion testing of thin-walled tubes. The stress-strain results from these tests followed two distinct trends: at low temperatures, strain hardening continued at shear strains of 8; at higher temperatures strain softening occurred. Continued strain hardening was observed in pure nickel, nickel-cobalt solid-solutions, pure aluminum, and two aluminum alloys. A laminar arrangement of closely spaced dislocation walls arises at large strains and low temperatures, which differs from the well-recovered equiaxed subgrain structure observed at high temperature. Thus, it appears that dynamic recovery processes are not sufficient to establish a steady-state dislocation structure at low temperatures. Strain softening in nickel and nickel-cobalt at higher temperatures was attributed to dynamic recrystallization. In none of the large strain tests conducted was a steady-state flow stress, independent of strain, observed. Torsion test data were found to match data from steady-state tensile creep and compression tests. In the case of the large strain torsion tests of nickel, recrystallization occurred. This suggests that recrystallization and boundary migration can be important processes in creep.  相似文献   

8.
The deformation microstructures of commercially pure aluminum deformed by plane strain compression to 50 pct thickenss reduction at temperatures between 100 °C and 300 °C, under two strain rates, 5×10−2 s−1 and 5×10−4 s−1, have been characterized by transmission electron microscopy. As the deformation temperature increases, the deformation microstructure gradually changes from a checkerboard pattern into an equiaxed subgrain structure with increasing subgrain size. The fraction of geometrically necessary boundaries (GNBs) found in warm-worked aluminum is much less than that found at room temperature. The average misorientation of dislocation boundaries appears to be independent of deformation temperature and strain rate. The constancy of the average misorientations is a combined effect of the variation of the fractions of GNBs and incidental dislocation boundaries (IDBs) and the variation of the average misorientations of GNBs and IDBs. Scaling theory can apply to both boundary misorientations and subgrain sizes that formed at different temperatures and strain rates. Subgrain size distributions for different temperatures and strain rates all resemble a lognormal distribution.  相似文献   

9.
The deformation microstructures of commercially pure aluminum deformed by plane strain compression to 50 pct thickness reduction at temperatures between 100 °C and 300 °C, under two strain rates, 5 × 10−2 s−1 and 5 × 10−4 s−1, have been characterized by transmission electron microscopy. As the deformation temperature increases, the deformation microstructure gradually changes from a checkerboard pattern into an equiaxed subgrain structure with increasing subgrain size. The fraction of geometrically necessary boundaries (GNBs) found in warm-worked aluminum is much less than that found at room temperature. The average misorientation of dislocation boundaries appears to be independent of deformation temperature and strain rate. The constancy of the average misorientations is a combined effect of the variation of the fractions of GNBs and incidental dislocation boundaries (IDBs) and the variation of the average misorientations of GNBs and IDBs. Scaling theory can apply to both boundary misorientations and subgrain sizes that formed at different temperatures and strain rates. Subgrain size distributions for different temperatures and strain rates all resemble a lognormal distribution.  相似文献   

10.
11.
The resistance to deformation of nitrogen-bearing Cr–Ni–Mn steel at 800–1200°C is investigated by means of the Gleeble 3800 system. By analysis of the deformation diagrams—in particular, determination of the threshold strain for dynamic recrystallization—the temperature and strain corresponding to the onset of dynamic recrystallization are established as a function of the strain rate, and optimal temperatures for hot stamping, forging, and rolling are recommended for industrial conditions. With true strain e = 0.9, the dynamic recrystallization in the steel at strain rates of 10–2–2 s–1 occurs at temperatures no lower than 900°C. Metallographic data confirm the experimental results and show that the structure formation in the steel on isothermal deformation at different rates is different above 900°C. With increase in temperature and decrease in strain rate, relaxation processes are more developed. At a strain rate of 0.01 s–1 (stamping on a press), dynamic recrystallization begins at around e = 0.1 (relative reduction around 10%) in the range 1100–1200°C. Strain of around 20 and 30%, respectively, is required with decrease in temperature to 1000 and 900°C. With increase in strain rate to 0.1 s–1 (forging), dynamic recrystallization begins with around 20% strain above 1100°C, 28% at 1000, and 35% at 900°C. At a strain rate of 1–2 s–1 (rolling), dynamic recrystallization begins at around 30% strain in the range 1000–1100°C. In that case, the threshold strain is 36% at both 900 and 1200°C.  相似文献   

12.
The behavior of a Ni-Cr-Co base alloy with significant additions of Mo, Ti and Al (Nimonic 105) under hot working conditions was studied using hot compression tests in the temperature range of 1223 to 1523 K and strain rates between 0.38 and 64.3 s-1. The microstructure of the Nimonic 105 is complex and the matrix contains second phases in the form of Ni3 (Ti, Al) dispersion (γ′), various Cr and Ti carbides and titanium cyanonitrides inclusions. However, the results show that above the dissolution temperature of the γ′ phase, the alloy behaves like a single phase nickel-base solid solution from the point view of steady state flow stress-temperature-strain rate relationships, and the activation energies for hot working and static recrystallization. Under deformation conditions where the γ′ phase is present, as in the case of creep, the activation energy is almost doubled. The hot working temperature range giving sound product is 1280 to 1450 K (170 K) at a strain rate of 0.4 s-1 and decreases to 1400 to 1480 K (80 K) at a strain rate of 65 s-1. At temperatures above the higher limit the alloy suffers intercrystalline cracks due to hot shortness and at temperatures below the lower limit the alloy suffers transcrystalline cracks due to excessive strain hardening.  相似文献   

13.
 Hot compression tests of microalloyed forging steels 38MnVS were carried out on the Gleeble-3800 thermo-mechanical simulator at the deformation temperatures from 950 to 1150 ℃ with the strain rates ranging from 0. 1 to 10 s-1. The effects of the deformation temperature and strain rate on the austenite dynamic recrystallization and microstructural changes were researched. The experimental results show that the dynamic recrystallization accelerated with the increase of the deformation temperatures and the decrease of the strain rate. The activation energy of dynamic recrystallization was calculated, which was about 275. 453 kJ/mol. The relation between the dynamic recrystallization and the Z-parameter was investigated, and the state chart of the dynamic recrystallization of the microalloyed forging steel 38MnVS was made according to the experimental data and the deformation parameters.  相似文献   

14.
Hot compression tests were conducted in a temperature range of 1173 K to 1323 K (900 °C to 1050 °C) and strain rates of 0.001 seconds−1 to 1 second−1 to investigate the hot deformation behavior of the austenitic stainless steel type 1.4563. The results showed that hot deformation at low temperatures, i.e., 1173 K to 1223 K (900 °C to 950°C), and at low and medium strain rates, i.e., 0.001 seconds−1 to 0.1 seconds−1, results in the dynamic formation of worm-like precipitates on existing grain boundaries. This in turn led to the restriction or even inhibition of dynamic recrystallization. However, at higher temperatures and strain rates when either the time frame for dynamic precipitation was too short or the driving force was low, dynamic recrystallization occurred readily. Furthermore, at low strain rates and high temperatures, there was no sign of particles, but the interactions between solute atoms and mobile dislocations made the flow curves serrated. The strain rate sensitivity was determined and found to change from 0.1 to 0.16 for a temperature increase from 1173 K to 1323 K (900 °C to 1050 °C). The variations of mean flow stress with strain rate and temperature were analyzed. The calculated apparent activation energy for the material was approximately 406 kJ/mol. The hyperbolic sine function correlated the Zener-Hollomon parameter and flow stress successfully at intermediate stress levels. However, at low levels of flow stress a power-law equation and at high stresses an exponential equation well fitted the experimental data.  相似文献   

15.
The hot workability of a near gamma titanium aluminide alloy, Ti-49.5Al-2.5Nb-1.1Mn, was assessed in both the cast and the wrought conditions through a series of tension tests conducted over a wide range of strain rates (10−4 to 100 s−1) and temperatures (850 °C to 1377 °C). Tensile flow curves for both materials exhibited sharp peaks at low strain levels followed by pronounced necking and flow localization at high strain levels. A phenomenological analysis of the strain rate and temperature dependence of the peak stress data yielded an average value of the strain rate sensitivity equal to 0.21 and an apparent activation energy of ∼411 kJ/mol. At low strain rates, the tensile ductility displayed a maximum at ∼ 1050 °C to 1150 °C, whereas at high strain rates, a sharp transition from a brittle behavior at low temperatures to a ductile behavior at high temperatures was noticed. Dynamic recrystallization of the gamma phase was the major softening mechanism controlling the growth and coalescence of cavities and wedge cracks in specimens deformed at strain rates of 10−4 to 10−2 s−1 and temperatures varying from 950 °C to 1250 °C. The dynamically recrystallized grain size followed a power-law relationship with the Zener-Hollomon parameter. Deformation at temperatures higher than 1270 °C led to the formation of randomly oriented alpha laths within the gamma grains at low strain levels followed by their reorientation and evolution into fibrous structures containing γ + α phases, resulting in excellent ductility even at high strain rates.  相似文献   

16.
Dynamic softening and substructural changes during hot deformation of a ferritic Fe-26Cr stainless steel were studied. The flow stress increased to reach a steady state in all the cases and the steady-state stress decreased with decreasing Z, the Zener-Hollomon parameter. A constant subgrain size was observed to correspond to the steady-state flow and the steady-state subgrain size increased with decreasing Z. Substructure examinations revealed that elongated, pancake-shaped subgrains formed in the early stage of deformation. Straight sub-boundaries and equiaxed subgrains developed progressively with strain, leading eventually to a stable substructure at strains greater than 0.7. During deformation at 1100 °C, dynamic recrystallization occurred by the migration and coalescence of subboundaries. Dynamic recovery dominated during deformation at 900 °C, resulting in the formation of fine equiaxed subgrains. Based on microstructural observations, the process of substructural changes during hot deformation was described by a schematic diagram.  相似文献   

17.
Torsion tests, on a 2014 + 0.13Zr alloy, were performed at temperatures in the range 573 to 773 K under strain rates ranging from 10−3 to 10 s−1. Transmission electron microscopy (TEM) inspection was performed in order to establish the role of the hot deformation on the hardening second-phase particles. The pinning effect of Al3Zr particles was also investigated. At the testing temperatures, the Al3Zr particles were stable, and no significant statistic changes, in terms of density and mean size, occurred during the tests. Small Al3Zr dispersoid particles inhibit recrystallization by pinning the grain and subgrain boundaries during hot deformation. Yet, they are particularly resistant to dislocation shear microstructure mechanism. Grains were elongated and contained a large number of sub-grains a few microns in width.  相似文献   

18.
Microstructure and texture evolution during the severe-plastic deformation (SPD) of unalloyed aluminum were investigated to establish the effect of processing route and purity level on grain refinement and subgrain formation. Two lots of aluminum with different purity levels (99.998 pct Al and 99 pct Al) were subjected to large plastic strains at room temperaturevia four different deformation processes: equal-channel angular extrusion (ECAE), sheet rolling, conventional conical-die extrusion, and uniaxial compression. Following deformation, microstructures and textures were determined using orientation-imaging microscopy. In commercial-purity aluminum, the various deformation routes yielded an ultrafine microstructure with a ∼1.5-μm grain size, deduced to have been formedvia a dynamic-recovery mechanism. For high-purity aluminum, on the other hand, the minimum grain size produced after the various routes was ∼20 μm; the high fraction of high-angle grain boundaries (HAGBs) and the absence of subgrains/deformation bands in the final microstructure suggested the occurrence of discontinuous static recrystallization following the large plastic deformation at room temperature. The microstructure differences were underscored by the mechanical properties following four ECAE passes. The yield strength of commercial-purity aluminum quadrupled, whereas the high-purity aluminum showed only a minor increase relative to the annealed condition.  相似文献   

19.
In the hot working of Al, the flow curves are usually monotonic, reaching saturation at a lower strain εs and stress σs as temperature rises and strain rate declines. Microstructural examination confirms that the dislocation density rises to a steady-state level through formation of an equiaxed subgrain substructure with constant dimension that is larger for lower stress. The energy dis-sipation efficiency estimated by dynamic materials modeling for flow curves of the above type is the result of dynamic recovery, not of dynamic recrystallization, which is characterized by flow curves with a peak and marked softening to a steady-state regime. Formerly with Concordia University  相似文献   

20.
The deformation microstructures of commercially pure aluminum deformed by plane strain compression to 50 pct thickenss reduction at temperatures between 100 °C and 300 °C, under two strain rates, 5×10−2 s−1 and 5×10−4 s−1, have been characterized by transmission electron microscopy. As the deformation temperature increases, the deformation microstructure gradually changes from a checkerboard pattern into an equiaxed subgrain structure with increasing subgrain size. The fraction of geometrically necessary boundaries (GNBs) found in warm-worked aluminum is much less than that found at room temperature. The average misorientation of dislocation boundaries appears to be independent of deformation temperature and strain rate. The constancy of the average misorientations is a combined effect of the variation of the fractions of GNBs and incidental dislocation boundaries (IDBs) and the variation of the average misorientations of GNBs and IDBs. Scaling theory can apply to both boundary misorientations and subgrain sizes that formed at different temperatures and strain rates. Subgrain size distributions for different temperatures and strain rates all resemble a lognormal distribution.  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