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 546 毫秒
1.
The increasing demand for reducing vehicle weight in the automotive and aerospace industries has raised the need to develop improved structural aluminum-based alloys. Thus, horizontal solidification experiment with the Al-7%Si-0.3%Mg (mass fraction) alloy was carried out. A water-cooled horizontal directional solidification device was developed and used. Microstructural characterization was carried out using traditional techniques of metallography, optical microscopy and SEM microscopy. The Thermo-Calc software was used to generate the solidification path of the investigated alloy with addition of 0.17% Fe (mass fraction). The effects of the thermal parameters such as the growth rate (VL), cooling rate (TC) and solidification local time (tSL) on the formation of the macrostructure and on the dendritic microstructure evolution were evaluated. A columnar to equiaxed transition (CET) was found for VL and TC values from 0.82 to 0.98 mm/s and from 1.71 to 2.55 °C/s, respectively. The microstructure was characterized by the measurement of the primary and secondary dendrite arm spacings (λ1 and λ2, respectively). Experimental laws of λ1 =f(VL, TC) and λ2 =f(tSL) were proposed. It is observed that the interdendritic region is composed of the following eutectic mixture: α(Al)+Si+π-Al8Mg3FeSi6+θ-Mg2Si.  相似文献   

2.
Horizontal directional solidification experiments were carried out with a monophasic Sn-2%Sb (mass fraction) alloy to analyze the influence of solidification thermal parameters on the morphology and length scale of the microstructure. Continuous temperature measurements were made during solidification at different positions along the length of the casting and these temperature data were used to determine solidification thermal parameters, including the growth rate (VL) and the cooling rate (TR). High cooling rate cells and dendrites are shown to characterize the microstructure in different regions of the casting, with a reverse dendrite-to-cell transition occurring for TR>5.0 K/s. Cellular (lc) and primary dendrite arm spacings (l1) are determined along the length of the directionally-solidified casting. Experimental growth laws relating lc and l1 to VL and TR are proposed, and a comparative analysis with results from a vertical upward directional solidification experiment is carried out. The influence of morphology and length scale of the microstructure on microhardness is also analyzed.  相似文献   

3.
Fretting wear of Fe41Co7Cr15Mo14C15B6Y2 bulk metallic glass (BMG) was investigated using the ball-on-disc measurement. The wear behaviors of the samples roughly follow the classical Archard wear law at room temperature. The influence of temperature on wear resistance (Rw) and microhardness (Hv) was evaluated. A good linear correlation was found between Rw and H for the Fe-based BMG. The results indicate that the hardness softening is responsible for the decrease of Rw at high temperature. The wear mechanism at high temperature was elucidated through the analysis of oxidation and debris morphology on the wear scar. Compared with traditional crystalline steels and other non-ferrous metal-based glassy alloys, the bulk amorphous steel shows higher wear resistance and hardness, which promotes its application as an advanced engineering material.  相似文献   

4.
对Ti--43Al--3Si (原子分数, %) 合金在3---100 μm/s 的生长速度下进行了系统的定向凝固实验. 研究了生长速度对固/液界面形态及初始过渡区组织演化规律的影响. 合金在3---60 μm/s 的生长速度范围内均以胞晶形态生长, 胞晶间距随着生长速度的增大而减小; 当生长速度达到90 μm/s 时, 开始出现枝晶生长. 在定向凝固初始启动阶段, 存在清晰的热过渡区, 热过渡区内Ti5Si3 相分布及过渡区组织与定向凝固区组织的关联性对于籽晶材料的引晶效果有重要影响. 生长速度在10 μm/s 以内时, 热过渡区内Ti5Si3 相分布连续, 且热过渡区组织与定向凝固区组织的关联性好, 有利于该合金的引晶.  相似文献   

5.
《Intermetallics》2007,15(10):1395-1400
Directional solidification (DS) of a binary Mg–4Al (wt.%) alloy was carried out to investigate the microstructures and microsegregation under controlled solidification conditions. In directional solidification the microstructure depends on the growth rate V because the cooling rate, which governs the solidification microstructure, is the product of the growth rate and the temperature gradient. The ability to produce simple and uniform microstructures in directional solidification enables us to correlate the formation of the microstructure and its characteristic length scales quantitatively with processing parameters. The morphology of the solid–liquid interface and the microstructure of both the mushy zone and the steady-state region were characterized at different levels of growth rates. With the help of an electron microprobe, microsegregation was determined in a specimen directionally solidified with cooling rates ranging from 0.06 to 0.8 K/s. The calculated microsegregation results based on the Scheil model deviated significantly from the experimental data, which is anticipated since back diffusion was not included due to the lack of diffusivity data.  相似文献   

6.
The solid-liquid interface morphology and solute segregation behaviour of AI-Li alloy 8090during unidirectional solidification were studied by the liquid metal quenehing method undervaried processing conditions.When solidification rate,RO.75 mm/min (temper-ature gradient,G_L=130℃/cm),the structure revealed of planar or dendritic interfacerespectively.With the increase of R,the interface morphology becomes cellular from planargradually,within a narrow range.And the greater the R,the,finer the dendrite.Segregationof element Cu and impurity elements Fe and Si are quite severe,the interface morphologymarkedly influences on solute segregation.During solidification at coarse dendrite interface,their segregation ratios are rather great and solidified structure is coarse.  相似文献   

7.
The composition of Al-Cu-Mn ternary eutectic alloy was chosen to be Al-32.5 wt.%Cu-0.6 wt.%Mn to the Al2 Cu and Al12 Cu Mn2 solid phases within an aluminum matrix(α-Al) from its melt. The Al-32.5 wt.%Cu-0.6 wt.%Mn alloy was directionally solidified at a constant temperature gradient(G=8.1 K·mm~(-1)) with different growth rates, 8.4 to 166.2 μm·s~(-1),by using a Bridgman-type furnace. The eutectic temperature(the melting point) of 547.85 °C for the Al-32.5 wt.%Cu-0.6 wt.%Mn alloy was obtained from the DTA curve of the temperature difference between the test sample and the inert reference sample versus temperature or time. The lamellar spacings(λ) were measured from transverse sections of the samples. The dependencies of lamellar spacings(λAl-Al2 Cu) and microhardness on growth rates were obtained as, λ_(Al-Al2Cu)=3.02 V~(-0.36), HV=153.2(V)~(0.035), HV=170.6(λ)~(-0.09) and HV=144.3+0.82(λ_(AlAl2 Cu))~(-0.50), HV=149.9+53.48 V~(0.25), respectively, for the Al-Cu-Mn eutectic alloy. The bulk growth rates were determined as λ~2_(Al-Al2 Cu)·V = 25.38 μm~3·s~(-1) by using the measured values of λ_(Al-Al2 Cu) and V. A comparison of present results was also made with the previous similar experimental results.  相似文献   

8.
Different from other alloys, the observation in this work on the dendritic mushy zone shows that the freckles are formed in two different regions before and after peritectic reaction in directional solidification of Sn–Ni peritectic alloys. In addition, the experimental results demonstrate that the dendritic morphology is influenced by the temperature gradient zone melting and Gibbs–Thomson effects. A new Rayleigh number (RaP) is proposed in consideration of both effects and peritectic reaction. The prediction of RaP confirms the freckle formation in two regions during peritectic solidification. Besides, heavier thermosolutal convection in samples with larger diameter is also demonstrated.  相似文献   

9.
Peng  Peng  Yue  Jin-mian  Zhang  An-qiao  Wang  Jia-tai  Fan  Jiang-lei 《中国铸造》2022,19(5):411-418

The morphologies of intermetallic phases (IMCs) during directional solidification of the Sn-Cu (L+Cu3Sn→Cu6Sn5) and Sn-Co (L+CoSn→CoSn2) peritectic systems were analyzed. The primary Cu3Sn and peritectic Cu6Sn5 phases in Sn-Cu alloy are IMCs whose solubility ranges are narrow, while both the primary CoSn and peritectic CoSn2 phases in Sn-Co alloy are IMCs whose solubility ranges are nil in equilibrium condition. The experimental results before acid corrosion shows that the dendritic morphology of both the Cu6Sn5 and CoSn2 phases can be observed. The investigation on the local dendritic morphology after deep acid corrosion shows that these dendrites are composed of small sub-structures with faceted feature. Faceted growth of the primary Cu3Sn and CoSn phases is also confirmed, and a faceted to non-faceted transition in their morphologies is observed with increasing growth velocities. Further analysis shows that the dendritic morphology is formed in the solidified phases whose solubility range is larger during peritectic solidification.

  相似文献   

10.
High wear resistance with a stable contact resistance is a prerequisite for electrical contact applications. Although Ag−CdO has been widely used as an electrical contact material, there are intrinsic problems of forming large Cd oxide particles and environmental regulations against using Cd. Newly developed Ag−SnO2 alloys are considered good candidates to replace Ag−CdO alloys due to their stable and fine oxide formation capabilities. In addition, further improvements in performance are expected in Ag−SnO2 alloys by alloy modification and/or solidification processing to produce finer and stable oxide dispersions through internal oxidation. The effect of the addition of Te and misch metal, which function as oxide forming elements, on Ag−Sn−In ternary alloy was investigated. Up to 0.5 wt.% of Te and misch metals were added and rapidly solidified to maximize their effect on fine oxide formation in an Ag matrix. Resulting microstructural changes and properties were evaluated through electron microscopy, spectroscopy, and hardness measurements. The role of Te addition was to provide nucleation sites for complex oxides such as In2TeO6 phase and to ensure fine and well dispersed SnO2 oxide particles. A rapid increase in size was observed for both grain and oxide particles when Te content exceeded 0.3 wt.%. Misch metal addition, on the other hand, had a pronounced effect on grain size reduction of the Ag matrix, and was interpreted as a consequence of decreasing the latent heat of solidification. Maximum hardness was achieved at 0.3 wt.% misch metal addition. In both cases, hardness decreased rapidly at 0.5 wt.% addition and was attributed to the large grain size of the matrix, and also large oxide particles aggregated in the matrix grains.  相似文献   

11.
Intermetallic Ti-45Al-8.5Nb-(W, B, Y) alloys were directionally solidified at constant growth rates (V) ranging from 10 to 400 μm/s under the temperature gradient G = 3.8 × 103 K/m. Quenching was performed at the end of directional solidification (DS) experiments. Microstructure evolution was investigated by analyzing the microstructures formed at the quenching interfaces and in the DS regions. The primary dendritic arm spacing (λ) decreases with increasing growth rate according to the relationship λ ∝ V−0.36. Both the width of columnar grain (λw) and the interlamellar spacing (λs) decrease with increasing growth rate according to the relationships λwV−1.13 and λs ∝ V−0.32, respectively. Lamellar microstructure initially disappears from the dendrites at the growth rate of 100 μm/s and subsequently from the interdendritic regions when the growth rate is up to 200 μm/s. The B2 particles can precipitate in the interdendritic regions.  相似文献   

12.
《Acta Materialia》2008,56(13):3146-3161
The effects of a high magnetic field on the cellular and dendritic morphology in the Al–Cu alloy during directional solidification have been investigated, and results show that morphological instability of cell and dendrite has occurred. Indeed, at lower growth speeds, a high magnetic field of 10 T caused the cell and dendrite to twist and deflect from the solidification direction. Regular tilted structure forms at moderate growth speeds and the secondary dendritic arm in the upstream direction is more developed than the one in the downstream direction. In the case where the primary trunk has not deflected from the solidification direction, the field has caused the side-branching and the tip-splitting of the cell. These experimental results may be attributed to the thermoelectric magnetic force in the solid cell and dendrite and the change of the surface chemical potential and surface tension of the cellular and dendritic tip.  相似文献   

13.
The composition of an Al–Cu–Mg ternary eutectic alloy was chosen to be Al–30 wt% Cu–6 wt % Mg to have the Al2Cu and Al2CuMg solid phases within an aluminum matrix (α-Al) after its solidification from the melt. The alloy Al–30 wt % Cu–6 wt % Mg was directionally solidified at a constant temperature gradient (G = 8.55 K/mm) with different growth rates V, from 9.43 to 173.3 μm/s, by using a Bridgman-type furnace. The lamellar eutectic spacings (λE) were measured from transverse sections of the samples. The functional dependencies of lamellar spacings λE (\({\lambda _{A{l_2}CuMg}}\) and \({\lambda _{A{l_2}Cu}}\) in μm), microhardness H V (in kg/mm2), tensile strength σT (in MPa), and electrical resistivity ρ (in Ω m) on the growth rate V (in μm/s) were obtained as \({\lambda _{A{l_2}CuMg}} = 3.05{V^{ - 0.31}}\), \({\lambda _{A{l_2}Cu}} = 6.35{V^{ - 0.35}}\), \({H_V} = 308.3{\left( V \right)^{ - 0.33}}\); σT= 408.6(V)0.14, and ρ = 28.82 × 10–8(V)0.11, respectively for the Al–Cu–Mg eutectic alloy. The bulk growth rates were determined as \(\lambda _{A{l_2}CuMg}^2V = 93.2\) and \(\lambda _{A{l_2}Cu}^2V = 195.76\) by using the measured values of \({\lambda _{A{l_2}CuMg}}\), \({\lambda _{A{l_2}Cu}}\) and V. A comparison of present results was also made with the previous similar experimental results.  相似文献   

14.
15.
The FeCoNiCrAl alloys have many potential applications in the fields of structural materials, but few attempts were made to characterize the directional solidification of high entropy alloys. In the present research, the microstructure and corrosion behavior of FeCoNiCrAl high entropy alloy have been investigated under directional solidification. The results show that with increasing solidification rate, the interface morphology of the alloy evolves from planar to cellular and dendritic. The electrochemical experiment results demonstrate that the corrosion products of both non-directionally and directionally solidified FeCoNiCrAl alloys appear as rectangular blocks in phases which Cr and Fe are enriched, while Al and Ni are depleted, suggesting that Al and Ni are dissolved into the NaCl solution. Comparison of the potentiodynamic polarization behaviors between the two differently solidified FeCoNiCrAl high entropy alloys in a 3.5%NaCl solution shows that the corrosion resistance of directionally solidified FeCoNiCrAl alloy is superior to that of the non-directionally solidified FeCoNiCrAl alloy.  相似文献   

16.
The comprehensive microstructural evolution of Mg–3, 6 and 9 wt.% Al alloys with respect to the solidification parameters such as thermal gradient (G), solidification velocity (V), cooling rate (G·V) and solute (Al) content were investigated in the present study. Various solidification techniques, including directional solidification, wedge casting, sand and graphite mould casting, gravity casting in a Cu mould and water quenching, were employed in order to obtain wide ranges of cooling rates between 0.05 and 1000 K s–1. The microstructural length scales of Mg–Al alloys, such as secondary dendrite arm spacing and primary dendrite arm spacing, were determined experimentally and compared with published models. In addition, the solidification parameters of morphological transitions such as cellular to columnar dendrite and columnar to equiaxed dendrite were also determined. Based on all the experimental data and the solidification model, a solidification map was built in order to provide guidelines for the as-cast microstructural features of Mg–Al alloys.  相似文献   

17.
The wear behavior of a weight fraction of particles with up to 30 wt.% in situ AlB2 flakes reinforced in Al-4Cu matrix alloy composites and fabricated by a squeeze casting method was investigated in a pin-on-disk abrasion test instrument against different SiC abrasives at room conditions. Wear tests were performed under the load of 10 N against SiC abrasive papers of 80, 100, and 120 mesh grits. The effects of sliding speed, AlB2 flake content, and abrasive grit sizes on the abrasive wear properties of the matrix alloy and composites have been evaluated. The main wear mechanisms were identified using an optical microscope. The results showed that in situ AlB2 flake reinforcement improved the abrasion resistance against all the abrasives used, and the abrasive wear resistance decreased with an increase in the sliding speed and the abrasive grit size. The wear resistances of the composites were considerably bigger than those of the matrix alloy and increased with increases in in situ AlB2 flake contents.  相似文献   

18.
Despite the excellent high temperature mechanical properties of the Ni3Al intermetallic compound, its application is still limited due to its inherently weak grain boundary. Recent research advances have demonstrated that the tensile ductility can be enhanced by controlling the grain morphology using a directional solidification. In this study, a series of directional solidification experiments were carried out to increase both the tensile ductility and the strength of Ni3Al alloys by arraying either the ductile phase of γ-Ni-rich dendrite fibers or the hard phase of β-NiAl dendrite fibers in the γ′-Ni3Al matrix. The dendrite arm spacing could be controlled by the solidification rate, and the volume fraction of the γ or β phase could be altered by the Al content, ranging from 23 at.% to 27 at.%. With an increasing Al content, the γ dendritic microstructure was transformed into the β dendrite in the γ′ matrix, thereby reducing the tensile ductility by increasing the volume fraction of brittle β dendrites in the γ′ matrix. With an increasing solidification rate, the dendrite arm spacing decreased and the tensile properties of Ni3Al varied in a complex manner. The microstructural evolution affecting the tensile behavior of directionally solidified Ni3Al alloy specimens with different solidification rates and Al contents is discussed.  相似文献   

19.
对γ/γ'--αMo定向共晶合金进行激光快凝实验. 结果表明, 激光快凝后,合金组织细化, 偏析减少, 硬度显著提高. 在激光脉冲平行于定向共晶生长方向时,熔凝区重新生长出更细密的定向共晶组织; 激光脉冲垂直于定向共晶生长方向时, 熔凝区由更致密的胞状晶和枝晶组成. 在熔凝区的枝晶组织中, 枝晶臂由γ固溶体及从该固溶体析出的细小γ'相组成; 枝晶间由含有γ'的γ固溶体和富Mo亚稳相的共晶组成, 亚稳相包括[αMo(ordered)], Ni3Mo和[Ni3Mo(ordered)].  相似文献   

20.
《Acta Materialia》2002,50(17):4275-4292
The microstructure, mechanical properties and weld hot cracking behaviour of a cast Al–Mg–Sc alloy containing 0.17 wt.% Sc were compared with those of a Sc-free alloy of similar chemical composition. Although this level of Sc addition did not cause grain refinement, the dendritic substructure appeared to be finer. There was a significant increase in the yield and tensile strength and the microhardness of the Al–Mg–Sc alloy relative to its Sc-free counterpart. A discontinuous precipitation reaction was observed at the dendritic cell boundaries. Microchemical analysis revealed segregation of Mg and Sc at these interdendritic regions. No improvement was observed in the resistance of the alloy to weld solidification cracking or heat affected zone (HAZ) liquation cracking. This is explained in terms of the inability of this level of Sc addition to refine the solidification structure and to influence the liquation of solute-enriched dendritic cell boundaries of the cast material.  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