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 625 毫秒
1.
2.
Lead, 17.1, 11.2, and 5 volume fraction copper (14, 9, and 4 wt pct Cu) alloys have been directionally solidified at constant growth velocities ranging from 1 to 100 μm s−1. Serially increasing the growth velocity within this range results in a graded microstructural transition from fully columnar, albeit segregated, copper dendrites in a lead matrix to one consisting only of equiaxed grains. The imposed velocity necessary to effect fully equiaxed growth is found to drop rapidly as the volume fraction of copper is decreased. Factors which complicate the controlled, directional solidification of these alloys are discussed and the experimental results are interpreted in view of, and seen to be in qualitative agreement with, Hunt’s theory on the transition from columnar to equiaxed growth of dendrites.  相似文献   

3.
Microstructures are shown for directionally solidified Al-3.4 wt pct Bi alloys with 0.2 wt pct Fe and 0.6 wt pct Fe additions. The third element causes the LI/SI + LII growth interface to become cellular. The bismuth forms at the cell nodes, appearing either as uniformly spaced arrays of spheres in the case of 0.2 wt pct Fe, or as an irregular net-work in the case of 0.6 wt pct Fe. Changes in the growth conditions which are known to control cellular structure are seen to have a similar effect on the bismuth spacing, with the cross sectional spacing varying as the inverse ofG (Temperature gradient) .R (Growth rate).  相似文献   

4.
To improve the high-temperature strength of Nb-Mo-Ti-Si in-situ composites, alloying with W and a directional solidification technique were employed. The alloy composition of Nb-xMo-10Ti-18Si (x=10 or 20) was used as the base, and Nb was further replaced by 0, 5, 10 and 15 mol pct W. For samples without W, the as-cast microstructure was a eutectic mixture of fine Nb solid solution (Nb SS ) and (Nb, Me)5 Si3 silicide (Me = Mo, W, or Ti), while large primary Nb SS particles appeared besides the eutectic mixture as a result of replacing Nb by W. The directionally solidified samples consisted of coarse Nb SS and (Nb,Me)5 Si3 silicides, and the microstructure showed a slight orientation in the direction of growth. The maximum compressive ductility (ɛ max) at room temperature decreased with increasing W content and was in the range of 0.8 to 2.3 pct, in contrast to the Vickers hardness (HV), which increased with W content. The 0.2 pct yield compressive strength (σ 0.2) and the specific 0.2 pct yield compressive strength (σ 0.2S ) (σ 0.2 divided by the density of sample) at elevated temperatures were markedly improved by the W addition. The directionally solidified samples always showed higher σ 0.2 and σ 0.2S values than the as-cast samples. At elevated temperatures, the directionally solidified sample with 10 mol pct Mo and 15 mol pct W had the highest σ 0.2 and σ 0.2S values; even at 1770 K, σ 0.2 was as high as 650 MPa. The directionally solidified materials alloyed with W exhibited excellent compressive creep performance. The sample with 10 mol pct Mo and 15 mol pct W had a minimum creep rate of 1.4×10−7s−1 and retained steady creep deformation at 1670 K and an initial stress of 200 MPa. Under compression, the damage and failure of these in-situ composites were dominated by decohesion of interfaces between the Nb SS and silicide matrix.  相似文献   

5.
Tin-lead alloys were solidified directionally and the position of the columnar to equiaxed transition determined on vertical sections of the ingots. The columnar length was found to increase with decreasing lead concentration and increasing heat transfer coefficient. A mathematical model of the heat flow in the system was used to determine local temperatures, temperature gradients, and velocities in the solidifying alloy. Comparing the position of the columnar to equiaxed transition and local thermal conditions, it was found that the transition occurred when the temperature gradient in the melt at the liquidus temperature was 0.11°C/mm for Sn 10 wt pct Pb, 0.10°C/mm for Sn 5 wt pct Pb, and 0.13°C/mm for Sn 15 wt pct Pb. The position of the transformation was found to be independent of melt superheat for the conditions considered.  相似文献   

6.
    
A series of Ni-Nb-Al-Cr(γ/γ′- σ) alloys in the composition ranges Nb 19.3 to 23.2 wt pct, Al 2.5 to 5.2 wt pct and Cr 0 to 7.05 wt pct have been directionally solidified under high thermal gradient (G) at both steady state and under conditions of abruptly or gradually changing growth rate(ft). The critical ratio of G andR, (g/r)*, to achieve two-phase plane frontin- situ composite growth increases as chromium and niobium (Cb) concentration deviates from the trough or surface of two-fold saturation. Interlamellar spacing of composites tend to decrease with increasing chromium content. Structures produced at steady state growth in whichG/R < (G/R)* are consistent with previous work and can be related to the location of the alloy composition with respect to the line of two-fold saturation. For alloys, which at lowG/R exhibited σ dendrites, any perturbation in growth velocity (atG/R > (G/R)*) precipitated a single phase σ (Ni3Nb) band. For alloys which at lowG/R exhibited γ dendrites a similar effect was achieved only when growth rate was reduced abruptly by more than an order of magnitude. Interlamellar spacing of two alloys (approximately Ni-20 wt pct Nb-2.5 wt pct Al-6 wt pct Cr) was studied and for abrupt reductions in growth rate in which bands were not produced, it was observed to decay slowly to the new steady state value over distances which are inconsistent with the assumption of simple niobium diffusion control. A gradual increase in growth velocity for one of these alloys resulted in extremely slow adjustment of interlamellar spacing occurring over a period greater than one hour. An abrupt increase in growth velocity for all alloys caused immediate adjustment of interlamellar spacing to the new steady state value. M. A. NEFF formerly a Graduate Student, Mass. Institute of Technology B. A. RICKINSON formerly Research Associate, Mass. Institute of Technology K. P. YOUNG formerly Research Associate, Mass. Institute of Technology  相似文献   

7.
Macrosegregation along the length of the directionally solidified samples is produced when Pb-Sn alloys (10 to 58 wt pct Sn) are directionally solidified in a positive thermal gradient (melt on top, solid below, and gravity pointing down) with steady-state dendritic arrayed morphology (the length of the mushy zone, much smaller than the initial length of the melt column, remaining nearly constant during growth). The extent of the macrosegregation increases with increasing tin content, becomes maximum for 33.3 wt pct Sn, and decreases with further increase in tin content. The intensity of the interdendritic thermosolutal convection responsible for the longitudinal macrosegregation can be represented by the effective partition coefficient(k E), anempirical parameter obtained from the dependence of the longitudinal macrosegregation on fraction distance solidified. The extent of the macrosegregation appears to be related to a parameter,x03BB; 2 1fE(CE−Ct)}, where A, is the primary dendrite spacing,f Eis the volume fraction of the interdendritic melt, andC EandC tare the eutectic composition and the melt composition ahead of the dendrite tips, respectively.  相似文献   

8.
The liquidus surface of the C-Cr-Fe system has been experimentally determined in the Fe-rich region —C ≤6 wt pct, Cr ≤40 wt pct —using a sensitive differential thermal analysis technique, along with optical and scanning electron microscopy and X-ray diffraction. Previous liquidus surfaces for this system have differed on the extent of the (Cr,Fe)23C6 liquidus field, with one version reporting its existence at ∼20 wt pet Cr, and others finding that it did not occur at Cr levels of less than ∼60 wt pct. The present investigation provides evidence in favor of the second contention, with the (Cr,Fe)23C6 field not being detected at Cr ≤40 wt pct. Changes are proposed to the accepted liquidus surface in respect of the compositions of the invariant reactions—L + αδFe ⇌γFe + (Cr,Fe)7C3 andL + (Cr,Fe)7C3γFe + (Fe,Cr)3C —and of the monovariant eutectic valley—L⇌ γFe + (Cr,Fe)7C3.  相似文献   

9.
The effects of electrochemically pre-dissolved hydrogen on room-temperature fracture initiation in Beta-C titanium (Ti-3Al-8V-6Cr-4Mo-4Zr wt pct) have been investigated using circumferentially notched tensile specimens. Finite element-based analysis of notch stress fields was used to define relationships between the local threshold stress for crack initiation vs total internal hydrogen concentration. The as-received, solution heat treated (ST, σy.2 pct=865 MPa) and the ST + peak-aged conditions (STA, σ y.2% pct=1260 MPa) were compared after defining the relationships between the fracture process zone hydrogen concentration, hydrogen-metal interactions (i.e., hydrostatic stress field occlusion, trapping, hydriding), and the resulting fracture initiation behavior of each. Solutionized + peak-aged (β+α) Beta-C fractured intergranularly above total hydrogen concentrations of ∼1000 wt ppm. (5.1 at. pct). A fracture mode consistent with cleavage occurred at ∼2100 wt ppm. (10.7 at. pct). Solutionized Beta-C resisted hydrogen-assisted cracking (e.g., did not crack intergranularly) but was not immune; cleavage cracking was provoked at ∼4000 wt ppm. (20.4 at. pct). Coldworked ST Beta-C (CW, σ y.2 pct=1107 MPa) did not crack intergranularly; fracture initiation behavior was similar to the ST condition regardless of specimen orientation. This suggests that high yield strength alone does not account for the susceptibility to intergranular cracking observed in the STA β+α condition. Stroke-rate studies and X-ray diffraction investigation of H partitioning suggests that equilibrium hydriding and/or irreversible trapping does not singularly control intergranular fracture initiation of the STA condition. Fractographic evidence and finite element results show that a finite plastic zone exists prior to intergranular fracture of the STA condition. This suggests that a criterion for fracture that incorporates plastic strain and stress should be considered.  相似文献   

10.
The quench sensitivity of cast Al-7 wt pct Si-0.4 wt pct Mg alloy   总被引:3,自引:0,他引:3  
The effect of quenching condition on the mechanical properties of an A356 (Al-7 wt pct Si-0.4 wt pct Mg) casting alloy has been studied using a combination of mechanical testing, differential scanning calorimetry (DSC), and transmission electron microscopy (TEM). As the quench rate decreases from 250 °C/s to 0.5 °C/s, the ultimate tensile strength (UTS) and yield strength decrease by approximately 27 and 33 pct, respectively. The ductility also decreases with decreasing quench rate. It appears that with the peak-aged condition, both the UTS and yield strength are a logarithmic function of the quench rate,i.e., UTS orσ y =A logR +B. The termA is a measure of quench sensitivity. For both UTS and yield strength of the peak-aged A356 alloy,A is approximately 32 to 33 MPa/log (°C/s). The peak-aged A356 alloy is more quench sensitive than the aluminum alloy 6063. For 6063,A is approximately 10 MPa/log (°C/s). The higher quench sensitivity of A356 is probably due to the high level of excess Si. A lower quench rate results in a lower level of solute supersaturation in the α-Al matrix and a decreased amount of excess Si in the matrix after quenching. Both of these mechanisms play important roles in causing the decrease in the strength of the peak-aged A356 with decreasing the quench rate.  相似文献   

11.
Directionally solidified Al-4 wt pct Cu-0.2 wt pct Ti alloy ingots with equiaxed structures were analyzed for solute distribution. Inverse segregation was found to exist, and the extent of segregation in the first 50 pct of the ingot solidified agreed well with that calculated from inverse segregation theory based on the flowback of enriched liquid. However, at higher points in the ingot, the solute distribution deviated from the theoretical one. Also, more and bigger shrinkage pores existed in the equiaxed grained ingot than in a columnar grained ingot. It is suggested that the formation of randomly distributed, equiaxed grains restricts the flowback of enriched interdendritic liquid much more than the formation of columnar grains where interdendritic channels exist. In an equiaxed grained ingot solidified vertically from the bottom upward, the hydrostatic pressure becomes insufficient to cause complete flowback of enriched liquid in the latter stages of solidification. This effect causes porosity in the ingot and a solute distribution which deviates from that predicted by the theory of inverse segregation.  相似文献   

12.
This study concerns the crystallographic identification and compositions of precipitates formed in superaustenitic stainless steel. Three experimental alloys, all containing 24 wt pct Cr, 22 wt pct Ni, and 0.5 wt pct N but with varying amounts of Mo and W, were investigated after sensitization heat treatment (aging) at 900 °C. The contents of Mo and W in the three alloys were 7 wt pct Mo, (6 wt pct Mo + 2 wt pct W) and (5 wt pct Mo + 5 wt pct W), respectively. While σ and x were the main secondary phases found in the W-free alloy, replacement of Mo by W was found to promote the formation of Laves-related phases with high Mo + W content. The complex crystallographic nature of Laves-related precipitates was exemplified through the formation of intergrowing C14 Laves, μ, and C phases, all with closely related crystal structures. There was no difference in chemical composition between the three phases. Prolonged aging resulted in intragranular precipitation of different intermetallic phases, as well as formation of nitrogen bearing phases, π and Cr2N, adjacent to previously formed intermetallic precipitates. The content of Mo + W was found to decrease with increasing aging time for all secondary phases.  相似文献   

13.
Combined additions of Ge and Si to Al are known to produce higher precipitation hardening than that which occurs in the constituent binaries, when the total amounts of alloying atoms are the same for all the alloys investigated. In the resultant Al-Ge-Si alloys, the diamond cubic precipitates contain both Ge and Si and are designated as GeSi. During artificial aging at 160 °C, the GeSi precipitates are commonly present in three forms, i.e., equiaxed, 〈100〉Al lath, and triangular plate. The equiaxed form is the dominant one of the three. This article examines the influence of varying amounts (i.e., 2 to 4 wt pct) of Cu additions on the morphology of GeSi precipitates formed in an Al-2.6 wt pct Ge-1.04 wt pct Si alloy during artificial aging at 160 °C. It is shown that Cu additions have the remarkable effect of maximizing the nucleation frequency of the 〈100〉Al lath form and simultaneously suppressing the nucleation of the equiaxed and the plate forms of the GeSi precipitates. Increasing Cu additions also increase the homogeneity and cause refinement of the 〈100〉Al laths. These results are discussed in light of (1) the critical requirement of vacancies for the nucleation and growth of GeSi precipitates having an atomic volume larger than Al and (2) the crystallographic nature of the negative dilation strains that develop locally in the Cu-rich regions of the Al matrix. It is further shown that, in the alloys containing increased levels (i.e., exceeding about 2.5 wt pct) of Cu, the precipitation of ϑ′ (metastable ϑ-Al2Cu) phase occurs, and that the nucleation of Cu-rich ϑ′ precipitates occurs upon the 〈100〉Al laths of GeSi. The latter effect is discussed in terms of the attainment of both the nucleation site and the necessary solute supersaturation at the 〈100〉Al GeSi/α-Al interfaces.  相似文献   

14.
The hot-working behavior of two metal matrix composites (7090 + 20 vol pct SiC whiskers and 6061 + 20 vol pct SiC whiskers) and their powder metallurgy matrix alloys (7090 and 6061) was studied by hot torsion testing. Flow stress (σo) and strain-to-failure (ε f ) data were generated at deformation temperatures and strain rates corresponding to the potential range for commercially hot-working these alloys. Based on the hot torsion data, hot-working parameters were recommended where σo was low and ε f was high. Strain rate sensitivities and activation energies of deformation were computed for the alloys. Formerly with Martin Marietta Laboratories, Baltimore, MD  相似文献   

15.
The columnar-to-equiaxed transition in Al 3 Pct Cu   总被引:1,自引:0,他引:1  
A columnar-to-equiaxed transition is observed in Al 3 pct Cu solidified directionally from a chill face. The transition occurs when the temperature gradient in the melt ahead of the columnar dendrites decreases to 0.6 ‡C/cm at dendrite growth rates of about 5 x 10-3 cm/s. Increasing the nuclei density by adding 171 ppm of TiB2 to the melt produces a fine-grained structure without columnar growth. Adding 100 ppm TiB2 has no effect on the cast structure or columnar-to-equiaxed transition. The results are considered in relation to the model for the columnar-to-equiaxed transition proposed by Hunt.[2]  相似文献   

16.
It has been reported that the mechanical properties and the corrosion resistance (CR) of metallic alloys depend strongly on the solidification microstructural arrangement. The correlation of corrosion behavior and mechanical properties with microstructure parameters can be very useful for planning solidification conditions in order to achieve a desired level of final properties. The aim of the present work is to investigate the influence of heat-transfer solidification variables on the microstructural array of both Al 9 wt pct Si and Zn 27 wt pct Al alloy castings and to develop correlations between the as-cast dendritic microstructure, CR, and tensile mechanical properties. Experimental results include transient metal/mold heat-transfer coefficient (h i), secondary dendrite arm spacing (λ2), corrosion potential (E Corr), corrosion rate (i Corr), polarization resistance (R 1), capacitances values (Z CPE), ultimate tensile strength (UTS, σ u ), yield strength (YS, σ y ), and elongation. It is shown that σ U decreases with increasing λ2 while the CR increases with increasing λ2, for both alloys experimentally examined. A combined plot of CR and σ U as a function of λ2 is proposed as a way to determine an optimum range of secondary dendrite arm spacing that provides good balance between both properties.  相似文献   

17.
The microstructure, tensile properties, and fractographic features of a near-α titanium alloy, IMI 829(Ti-6.1 wt pct Al-3.2 wt pct Zr-3.3 wt pct Sn-1 wt pct Nb-05 wt pct Mo-0.32 wt pct Si) have been studied after aging over a temperature range of 550°C to 950°C for 24 hours following solution treatment in the β phase field at 1050°C and water quenching. Transmission electron microscopy studies revealed that aging at 625°C and above produced discrete silicides at α′ interplatelet boundaries. However, aging at 900°C and above has also resulted in the precipitation of β phase along the lath boundaries of martensite. The silicides have been found to have a hexagonal structure withc=0.36 nm anda=0.70 nm (designated as S2 by earlier workers). There is a significant improvement in yield and ultimate tensile strength after aging at 625°C, but there is less improvement at higher aging temperatures. The tensile ductility is found to be drastically reduced. While the fracture surface of the unaged specimen shows elongated dimples, the aged samples show a mixed mode of fracture, consisting of facets, featureless parallel bands, and extremely fine dimples.  相似文献   

18.
γ/γ/’-δ eutectic alloy containing 21.5 wt pct Nb, 2.5 wt pct Al, balance Ni was directionally solidified under a thermal gradient (G) of 500 K/cm and different schedules of growth rate (R). Under steady state growth rate conditions, the critical (G/R) for plane front solidification is 25 K h/cm2. At progressively lower (G/R) the structure becomes richer in γ-phase,i.e. hypoeutectic. Sudden increase in growth speed causes the structure to change from lamellar to cellular and gradual increase in growth speed results in interlamellar spacings that are larger than the extremum values. Sudden decrease in growth speed causes little disturbance in the structure but causes the eutectic grain size to increase.  相似文献   

19.
The morphological and kinetic nature of corrosion of directionally solidified aluminum-4.5 wt pct copper alloy in the as-cast, solutionized and solutionized-and-aged conditions in air-saturated aqueous 3.5 wt pct NaCl solution were evaluated. In the solutionized and solutionized-and-aged conditions the intergranular attack and pitting are similar to those occurring in solutionized wrought alloys; the extent of attack at long times increases with increasing severity of solidification rate. The as-cast alloy exhibits a cored dendritic structure with significant formation of interdendritic nonequilibrium eutectic. Extensive inter dendritic corrosion of the α-phase containing more than 3.2 wt pct copper is seen; α containing less than 3.2 wt pct copper and the θ-Al2Cu phase are cathodic. Corrosion of the as-cast alloy is parabolic with time and increases with increasing severity of solidification rate in proportion to the amount of nonequilibrium second phase.  相似文献   

20.
The microstructures of Al-3Ti-lCe (wt pct) and Al-5Ti-5Ce alloys melt-spun under controlled He atmosphere have been characterized using analytical electron microscopy. The rapidly solidified microstructures comprise uniform, fine-scale dispersions of intermetallic phase in an aluminum matrix, and particular attention has been given to identification of the dispersed phases. In the Al-3Ti-lCe alloy, the dispersed particles are polycrystalline with a complex twinned substructure and a diamond cubic crystal structure(a o =1.44 ±0.01 nm) and composition consistent with the ternary compound Al20Ti2Ce (Al18Cr2Mg3 structure type, space group Fd3m). In the Al-5Ti-5Ce alloy, there is, in addition to the dispersed ternary phase, a separate uniform array of fine-scale particles of the binary compound Al11Ce3. The majority of such particles have the body-centered orthorhombic structure of the low-temperature polymorph, α-Al11Ce3, but there is evidence to suggest that at least some particles developvia initial formation of the high-temperature body-centered tetragonal phase, β-Al11Ce3. The accumulated evidence suggests that both binary and ternary particles formed as primary phases directly from the melt during rapid solidification, leaving only small concentrations of solute in aluminum matrix solid solution. Both phases are observed to be resistant to coarsening for up to 240 hours at 400 °C. Formerly Research Fellow, Department of Materials Engineering, Monash University.  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