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 109 毫秒
1.
The deformation behavior of a Burgers oriented α-β-Ti-13Mn bimetallic bicrystal was studied at two plastic strains, 0.52 and 2.08 pct. Two single crystals, α and β, each corresponding to the orientation of its respective bicrystal component were also investigated. The stress axes were and [1218]β. The interface planes were and and lay in the x’-z’ plane. The deformation behavior of the a component differed from that of the a single crystal because of plastically induced stresses,T y’z’ ,T x’z’ ,T x’y’ , and σ x’x’ . Prismatic slip and twinning were found in the single crystal α whereas the bicrystal revealed additionally pyramidal andc + a slip, the latter at the interface. Slip on the front and back surfaces was different and both thec + a and twinning systems acted to maintain compatibility. Slip in the β single crystal and the β bicrystal component were quite similar. However, there were differences in the intensity and amount of primary slip, (231) , on the front and back surfaces. The diminished amount of (231) slip on the back surface was due to plastically induced stresses, and on the front surface the primary slip cross slipped to slip which triggeredc + a slip in α. On the back surface the dominant slip system was which acted in response to the plastically induced stresses. An approximate calculation revealed that the interface deformation zone had about twice the flow stress of the average bicrystal stress. Formerly a Graduate Student in the Department of Physical and Engineering Metallurgy at Polytechnic Institute of New York, Brooklyn, NY  相似文献   

2.
Diffusion studies were carried out in the Fe-Ni-Al system at 1000 °C with solid-solid diffusion couples assembled with β (B2), β′ (bcc), and γ (fcc) single-phase alloys for the development of diffusion structures, diffusion paths, and for the determination of interdiffusion and intrinsic diffusion coefficients. The diffusion structures were examined by optical and scanning electron microscopy, and the concentration profiles were determined by electron microprobe analysis. Diffusion couples included several series of β vs γ and β′ vs γ diffusion couples characterized by a common terminal alloy bonded to several terminal alloys with varying compositions. The development of planar and nonplanar interfaces, as well as two-phase layers, as observed in various couples, were related to the diffusion paths. The interdiffusion fluxes of individual components were calculated directly from the experimental concentration profiles, and the diffusional interactions among components were examined in the light of zero-flux planes (ZFPs) and flux reversals, which were identified in several couples. Ternary interdiffusion coefficients ( (i, j = Al, Ni)), with Fe considered as the dependent concentration variable, were evaluated at composition points of the intersection of diffusion paths of single-phase couples and of multiphase couples that developed planar interfaces. The interdiffusion coefficients were the largest in magnitude for the β′ alloys, especially near the β/β′ miscibility gap, and decreased for the β and γ alloys. In the β and γ phases, the main interdiffusion coefficient for Al was larger than those for Ni and Fe. Also, Fe interdiffused faster than Ni in the Fe-rich β and β′ phases. The cross-interdiffusion coefficients ( and ) were negative in all three phases. In general, the coefficients were larger in magnitude than the coefficients; however, the magnitude of was greater than that of near the β/(β + γ) phase boundary on the ternary isotherm. In the β phase, the magnitude of (i, j=Al, Ni) coefficients increased over 1 to 2 orders of magnitude with a decrease in the Al concentration and increase in the Fe/Ni concentration ratio. Interdiffusion coefficients, extrapolated from the ternary coefficients for binary alloys, were consistent with those in literature. Intrinsic diffusion coefficients were also determined at selected compositions. In addition, tracer diffusion coefficients were estimated for the binary Fe-Al and Ni-Al alloys at selected compositions, from an extrapolation of ternary interdiffusion coefficients.  相似文献   

3.
A Ti-24Al-11Nb alloy has been heat-treated so as to obtain a microstructure of coarse α2 particles (D019 structure based on Ti3Al) in a matrix of the ordered βo phase (B2 structure based on Ti2AlNb). Dislocation structures generated by tensile strains of ∼2 pct at room temperature have been analyzed by transmission electron microscopy The βo phase is shown to deform inhomogeneously on {110}, {112}, and {123} planes by α/〈211〉 slip. The slipband structure is complex, consisting of segments of heavily pinned edge dislocations with periodic cross slip of screw components on to secondary slip planes. Incompatibility stresses at α2o interfaces can generate fine α[100] slip as well. The α2 phase deforms independently by α dislocation slip. Slipbands in the βo phase can shear the α2 phase by activatingc +a/2 slip on and slip planes, as well asa slip on higher order pyramidal planes, where the parallelism of the specific slip system is permitted by the Burgers relationship between the two phases.  相似文献   

4.
The microstructure of a casting Zn-40 wt pct Al alloy was studied by transmission electron microscopy (TEM). The results show that primary α’ dendrites decompose into η-Zn phases and an aluminum matrix during cooling and at ambient aging. There are three typical morphologies corresponding to the different microcompositions or cooling conditions: coherent platelike η-Zn phases, spherical η-Zn phases coexisting with the platelike η-Zn phases, and the full equilibrium structure of spherical η-Zn phases and an aluminum matrix. The formation mechanism of these three kinds of microstructures is discussed. On the surface of the α’ dendrite, there is a cell with fcc structure, which should be β-Zn phase by the peritectic reaction. In the interdendritic regions, there are proeutectic β-Zn phases with equiaxed grains and lamellar colonies which are ternary eutectic β-Zn, η-Zn, and ε-CuZn4 phases, with the orientation relationships of
The T’ or T phase (Al4ZnCu3), which has an ordered bcc structure (a=0.296 nm), was found.  相似文献   

5.
A novel hexagonal phase (designated H(2)) has been detected as a major constituent of both Al62Ti10V28 and Al55Ti10V35 alloys, following chill casting and after homogenization at 1523 K. The phase H2 has an ordered hexagonal crystal structure (space group P63/mmc, α=0.558±0.001 and c=0.450±0.001 nm), similar to that of α 2-Ti3Al, and an atomic composition of 54±1 Al−11±1 Ti−35±1 V in the chill-cast Al55Ti10V35 alloy. A fine-scale, duplex lamellar structure, developed within the ordered H2 phase in the solid state, was composed of parallel-sided multivariants of ξ-Ti5Al11 phase, formed parallel to (0001)H 2. The orientation relationship between constituent phases was of the form
Following homogenization (1 hour at 1523 K) and isothermal heat treatment at temperatures in the range from 1073 to 1273 K, the Al62Ti10V28 and Al55Ti10V35 alloys were observed to develop two-phase microstructures comprising β-(Ti, V)/δ-Al3(Ti,V) and β/ξ-Ti5Al11, respectively. In both alloys, the β phase could be retained as a stable phase to room temperature, regardless of cooling rate, whereas the H2 phase was unstable at temperatures below ∼1273 K and was consumed by the formation of ξ phase at lower temperatures.  相似文献   

6.
The competition of slip among , and slip modes of hcp metals has been analyzed geometrically in terms of a critical resolved shear stress, CRSS, criterion. Under the action of an applied stress slip systems of one or more modes may be activated depending on the value of the CRSS and on the orientation of the slip systems with respect to the applied stress. If the CRSS of a given slip mode should exceed a limiting value relative to the CRSS of the other modes, however, the given mode becomes inoperative even under the most favorably stressed conditions. It is found by an examination of the yield loci that basal slip is inoperative if α2 < cos θ; prism slip is inoperative if α2 < α1 sin θ; and pyramidal slip is inoperative if α2 > cos θ + α1 sin θ where and are, respectively, the ratios of CRSS for prism and pyramidal slips relative to basal slip, and ϕ is the angle between the (0001) and normals. Since the value of ϕ is a function ofc/a, the limiting values of α1 and α1 depend on thec/a ratio of the crystal structure.  相似文献   

7.
The ductility of Mg alloys is limited due to a shortage of independent slip systems. In particular, c-axis compression cannot be accomodated by any of the easy slip or twinning modes. Basaltextured samples of pure Mg and Mg-15 at. pct Li were examined for the presence of <c+a> dislocations by post-mortem transmission electron microscopy (TEM) after a small deformation, which forced the majority of grains to compress nearly parallel to their c-axes. A higher density and more uniform distribution of <c+a> dislocations is found in the Li-containing alloy. Because the pyramidal slip mode coffers five independent slip systems, it provides a satisfying explanation for the enhanced ductility of α-solid solution Mg-Li alloys as compared to pure Mg. The issue of <c+a> dislocation dissocation and decomposition remains open from an experimental point of view. Theoretically, the most feasible configuration is a collinear dissociation into two 1/2 <c+a> partial dislocations, with an intervening stacking fault on the glide plane. It is speculated that Li additions may lower the fault’s energy and, thereby, increase the stability of this glissile configuration. This article is based on a presentation made in the symposium entitled “Defect Properties and Mechanical Behavior of HCP Metals and Alloys” at the TMS Annual Meeting, February 11–15, 2001, in New Orleans, Louisiana, under the auspices of the following ASM committees: Materials Science Critical Technology Sector, Structural Materials Division, Electronic, Magnetic & Photonic Materials Division, Chemistry & Physics of Materials Committe, Joint Nuclear Materials Committee, and Titanium Committee.  相似文献   

8.
We calculated in great detail all the relative partial molar thermodynamic quantities of Bi, , and , where Q=H, S, and G, of the molten Bi-Sn system from their experimental counterparts, and Δ , using the authentic Chiotti integration and assessed their unequivocal limiting values at infinite dilution of both components, and , by the genuine Chiotti derivative. Thermodynamic consistency tests were satisfied. While the excess partial quantities, , and , oscillate with composition, their excess integral quantities, ΔH m, Δ i S m xs , and ΔG m xs , are rather bland; and the integral quantities, ΔS m and ΔG m, are nearly symmetrical with composition. The Bi-Sn solution is definitely nonregular and the solution reaction is exothermic (ΔH m<0). Atomistically, the solution has a weak segregation tendency, shown by the values in the relation TS m xs |>ΔG m xs >|ΔH m|. Although the excess integral molar heat capacity of our solution is zero, ΔC p,m xs =0, the solution does not belong to the Guggenheim’s athermal type by definition.  相似文献   

9.
A new experimental method is given for recovering the probability-distribution function S v ( |Δg). The function S v ( |Δg) is the grain-boundary area per unit volume as a function of grain-boundary plane orientation ( ), given a lattice misorientation (Δg) between the adjoining grains. The grain-boundary normal ( ) is expressed in the crystal frame in which the misorientation Δg originates. The proposed method recovers the three-dimensional S v ( |Δg) function using data taken from two-dimensional section planes. The method requires the measurement of many grain-boundary trace (in-plane) angles and lengths associated with grain boundaries of lattice misorientation. All such boundary traces may be observed from a single section plane if the crystallographic texture is sufficiently random. In heavily textured microstructures, the method requires the researcher to observe traces from multiple oblique section planes cut through the material. A method of quantitatively estimating whether the texture is sufficiently random is given. Simulations on both textured and nontextured microstructures demonstrate the validity of the method. Experimentally, the new method is used to analyze boundaries of misorientation (Σ3) observed in 304 stainless steel. Calculated grain-boundary plane-probability functions are shown to be consistent with what is already known. This article is based on a presentation made at the symposium “Characterization and Representation of Material Microstructures in 3-D” held October 8–10, 2002, in Columbus, OH, under the auspices of ASM International’s Phase Transformations committee.  相似文献   

10.
The core structures of 〈c+a〉 dislocations in hexagonal-close-packed (hcp) metals have been investigated by molecular dynamics (MD) simulation using a Lennard-Jones-type pair potential. The 〈c+a〉 edge dislocation has two types of core at 0 K; one is a perfect dislocation (type A), and the other has two 1/2 〈c+a〉 partials (type B). Type A transforms to type B by abruptly increasing temperature from 0 K to 293 K, while type B is stable in temperature range from 0 K to 293 K. In contrast, type A extends parallel to (0001) at 30 K, and this extended core is still stable at 293 K. These results suggest that the 〈c+a〉 edge dislocation glides on the as two 1/2 〈c+a〉 partial dislocations and becomes sessile, due to changes of the core structure. The 〈c+a〉 screw dislocation spreads over two planes at 0 K. The core transforms into a unsymmetrical structure at 293 K, which is spread over and , and core spreading occurs parallel to at 1000 K. A critical strain to move screw dislocations depends on the sense of shear strain. The dependence of the yield stress on the shear direction can be explained in terms of these core structures. This article is based on a presentation made in the symposium entitled “Dect Properties and Mechanical Behavior of HCP Metals and Alloys” at the TMS Annual Meeting, February 11–15, 2001, in New Orleans, Louisiana, under the auspices of the following ASM committees: Materials Science Critical Technology Sector, Structural Materials Division, Electronic, Magnetic & Photonic Materials Division, Chemistry & Physics of Materials Committee, Joint Nuclear Materials Committee, and Titanium Committee.  相似文献   

11.
General expressions are derived for the calculation of the total product phase grain edge length per unit volume L V βββ and the total number of product phase quadruple points per unit volume QV (i.e., ββββ quadruple points) at any given time during solid state transformations occurring by nucleation and growth process. It is shown that, {
  相似文献   

12.
In the hot working at constant strain rate ( ) of Al and α Fe alloys at 0.5 to 0.9 T M (absolute melting temperature), steady-state deformation is achieved in similarity to creep, which is usually at constant stress. After an initial strain-hardening transient, the flow stress becomes constant in association with a substructure which remainsequiaxed and constant in the spacing of sub-boundaries and of dislocations in both walls and subgrains. All these spacings become larger at higher temperature (T) and lower values as well as with lower stress, being fully consistent with the relationships established in creep. Because hot working can proceed to a much higher true strain in torsion (∼100) and compression (∼2) as well as in extrusion (∼20) and rolling (∼5), it is possible to confirm that grains continue to elongate while the subgrains within them remain equiaxed and constant in size. When the thickness of grains reaches about 2 subgrain diameters (d s), the grain boundaries with serrations (∼d s) begin to impinge and the grains pinch off, becoming somewhat indistinguishable from the subgrains; this has been called geometric dynamic recrystallization (DRX). In polycrystals as at 20 °C, deformation bands form and rotate during hot working according to the Taylor theory, developing textures very similar to those in cold working. In metals of lower dynamic recoverability such as Cu, Ni, and γ Fe, new grains nucleate and grow (discontinuous DRX), leading to a steady state related to frequently renewed equiaxed grains, containing an equiaxed substructure that develops to a constant character and defines the flow stress. This article is based on a presentation made in the workshop entitled “Mechanisms of Elevated Temperature Plasticity and Fracture,” which was held June 27–29, 2001, in San Diego, CA, concurrent with the 2001 Joint Applied Mechanics and Materials Summer Conference. The workshop was sponsored by Basic Energy Sciences of the United States Department of Energy.  相似文献   

13.
Phase transformation of Zn-4Al-3Cu alloy during heat treatment   总被引:2,自引:0,他引:2  
The phase transformation in Zn-4 Al-3 Cu alloy employing various solution-treatment temperatures (230 °C to 325 °C) was studied by means of microhardness, scanning electron microscopy (SEM), electron probe microanalysis (EPMA), transmission electron microscopy (TEM), and X-ray diffraction (XRD). The starting microstructure of the as-cast Zn-4Al-3Cu alloy consists of an α phase (aluminum-rich, fcc structure) in the η matrix (zinc-rich, h.c.p. structure) prior to solution-treatment. A platelike ε phase with 3-μm length and 0.5-μm thickness was found in the η phase matrix after solution-treating the as-cast material at 240 °C for 1 hour. The ε phase was then dissolved gradually back into the η matrix above that temperature. A four-phase transformation, α + εT′ + η, was observed from the temperature 250 °C to 310 °C, wherein the T′ phase formed at the interface of ε platelet and η phase matrix. This T′ phase was further identified as a rhombohedral structure. As the solution-treatment temperature was increased to above 310 °C, the ε phase was completely dissolved back into the η matrix and numerous β phase particles were distributed uniformly in the η matrix. The β phase subsequently decomposed at room temperature to a fine α phase embedded in the η matrix. For the materials solution-treated above 250 °C, the microhardness of the η matrix increased in 40 minutes during natural aging, which was associated with the formation of fine ε phase of 0.15-μm diameter. The orientation relationship between this fine ε phase and η phase was determined as .  相似文献   

14.
Ternary interdiffusion coefficients were determined at 1000°C at several Fe−Ni−Al alloy compositions with multiphase β(bcc)vs γ (fcc) diffusion couples which developed planar β/γ-interfaces. The coefficients, (i,j=Al or Ni) were calculated at compositions corresponding to points of intersections of diffusion paths with Fe taken as the dependent component. These coefficients varied with composition by 1 to 2 orders of magnitude in the β-phase but relatively little in the γ-phase. Empirical relations were derived to describe the composition dependence of the main coefficients. and . Interdiffusion coefficients with either Al or Ni as the dependent component were also evaluated. The relative diffusivities of the elements increase in the order, Fe, Ni, Al for both β- and γ-alloys. The ternary diffusion data were consistent with binary interdiffusion coefficients for Fe−Al and Fe−Ni alloys. G. H. CHENG, formerly a Graduate Student at Purdue University  相似文献   

15.
16.
The enthalpies of formation of liquid (Cu + Mn) alloys were measured in the isoperibolic heat-flux calorimeter at 1573 K in the entire range of compositions. The integral molar enthalpy of mixing was found to be negative in the range of molar fractions 0 < x Mn < 0.31, with ΔH(min) = −0.69 ± 0.27 kJ mol−1 at x Mn = 0.12, and positive in the range 0.31 < x Mn < 1, with ΔH(max) = 3.67 ± 0.36 kJ mol−1 at x Mn = 0.75. Limiting partial molar enthalpies of manganese and copper were calculated as = −18.0 ± 6.6 kJ mol−1 and = 29.1 ± 4.9 kJ mol−1, respectively. The results are discussed in comparison with the thermodynamic data available in the literature and the equilibrium phase diagram.  相似文献   

17.
Texture evolution inα-Zr due to uniaxial deformation at 923 to 1123 K was investigated in crystal-bar Zr and Zr-2.5Nb. The temperature range selected corresponds to the two-phase (α +β) field in the Zr-2.5Nb alloy. It was found that uniaxial compression causes a progressive rotation of the (0002) plane normals away from the compression direction and away from the compression plane. In the crystal-bar Zr, the compression texture consists of a [0001] fiber tilted 30 deg from the compression axis. By contrast, in Zr-2.5Nb, a [0001] fiber with an angular spread of 30 deg is obtained. The effect of theβ phase present in Zr-2.5Nb at the temperatures investigated was evaluated by testing a Zr-20Nb alloy in compression. The β-phase texture consisted of a weak 〈111〉-〈00l〉 double fiber. Comparison of this texture and the textures observed in Zr-2.5Nb indicates that theβα transformation takes place by the growth of pre-existing a grains and not according to the Burgers mechanism. This transformation has, therefore, no direct effect on the α-phase texture after cooling to room temperature from the (α +β) field. Uniaxial elongation by swaging of Zr-2.5Nb produces a dual fiber. Similar results are obtained in hot extruded rods. Modeling of the development of textures in the α phase was performed using linear programming and employing relaxed constraint (RC) models (“curling” for tension and ”pancake” for compression) implemented for hexagonal close-packed (hcp) grains. It is assumed that prismatic, basal, and 〈c +a〉 pyramidal slip were the active deformation modes at high temperatures. It is shown that these models reduce the activity of the pyramidal slip systems to realistic values, in contrast to the full constraint (FC) approach, where most of the deformation is accommodated by 〈c +a〉 slip. Microstructural evidence is presented regarding the occurrence of ”curling” during uniaxial elongation. Formerly Graduate Student with the Department of Metallurgical Engineering, McGill University  相似文献   

18.
The evolution of microstructure and texture in the AZ-series Mg alloys subjected to electron-beam welding and gas tungsten arc welding are examined. Electron-beam welding is demonstrated to be a promising means of welding delicate Mg plates, bars, or tubes with a thickness of up to 50 mm; gas tungsten arc welding is limited to lower-end thin Mg sheets. The grains in the fusion zone (FZ) are nearly equiaxed in shape and ∼8 μm or less in size, due to the rapid cooling rate. The as-welded FZ microhardness and tensile strength are higher than the base metals due to the smaller grain size. The weld efficiency, defined as the postweld microhardness or tensile strength at the mid-FZ region divided by that of the unwelded base metal, is around 110 to 125 pct for electron-beam welding and 97 to 110 pct for gas tungsten arc welding. There are three main texture components present in the electron-beam-welded (EBW) FZ, i.e., (with TD// ), (with ND ∼15 deg), and (with WD ∼30 deg), where TD, ND, and WD are the transverse, normal, and welding directions, respectively. The crystal growth tends to align toward the most closed-packed direction, . The texture in gas tungsten arc welded (GTAW) specimens is more diverse and complicated than the EBW counterparts, due to the limited and shallow FZ and the lower cooling rate. The cooling rates calculated by the three-dimensional (3-D) and two-dimensional (2-D) heat-transfer models are considered to be the lower and upper bounds. The cooling rate increases with decreasing Al content, increasing weld speed, and increasing distance from the weld top surface. The influences of the FZ location, welding speed, and alloy content on the resulting texture components are rationalized and discussed.  相似文献   

19.
The ductility of Mg alloys is limited due to a shortage of independent slip systems. In particular, c-axis compression cannot be accommodated by any of the easy slip or twinning modes. Basal-textured samples of pure Mg and Mg-15 at. pct Li were examined for the presence of 〈c+a〉 dislocations by post-mortem transmission electron microscopy (TEM) after a small deformation, which forced the majority of grains to compress nearly parallel to their c-axes. A higher density and more uniform distribution of 〈c+a〉 dislocations is found in the Li-containing alloy. Because the 1/3〈11 3〉 {11 } pyramidal slip mode offers five independent slip systems, it provides a satisfying explanation for the enhanced ductility of α-solid solution Mg-Li alloys as compared to pure Mg. The issue of 〈c+a〉 dislocation dissociation and decomposition remains open from an experimental point of view. Theoretically, the most feasible configuration is a collinear dissociation into two 1/2〈c+a〉 partial dislocations, with an intervening stacking fault on the glide plane. It is speculated that Li additions may lower the fault’s energy and, thereby, increase the stability of this glissile configuration. This article is based on a presentation made in the symposium entitled “Defect Properties and Mechanical Behavior of HCP Metals and Alloys” at the TMS Annual Meeting, February 11–15, 2001, in New Orleans, Louisiana, under the auspices of the following ASM committees: Materials Science Critical Technology Sector, Structural Materials Division, Electronic, Magnetic & Photonic Materials Division, Chemistry & Physics of Materials Committee, Joint Nuclear Materials Committee, and Titanium Committee.  相似文献   

20.
M2C carbides were observed to precipitate within the austenite matrix of an Fe-24.6Mn-6.6Al-3. IMo-1.0C alloy after quenching from 1200 °C and aging at 750 °C. By means of transmission electron microscopy (TEM) and diffraction techniques, the orientation relationships between M2C (p) and the austenite (γ) matrix were determined to be: (0001)p//(111)γ, (11– 0)p// (1 0)γ, ( 100)p//(11 )γ. M2C carbide has been reported by many researchers to precipitate from the ferrite matrix or along austenite/ferrite boundaries in alloy steels containing Mo. However, little information concerning the formation of M2C in the austenite matrix has been provided. This investigation presents the first evidence for the existence of M2C carbide wholly within the austenite matrix and its relationship to the austenite. The energy-dispersive spectrometry (EDS) analyses were performed on M2C carbides, and the results indicate that the solubility of the M2C carbide for foreign atoms other than Mo is very limited.  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