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 15 毫秒
1.
Su  Yunpeng  Lin  Xin  Wang  Meng  Huang  Weidong 《Journal of Materials Science》2021,56(25):14314-14332

Peritectic solidification has attracted increasing attention as a lot of important binary alloys, such as Fe–Ni, Zn–Cu, Fe–C and Ti–Al, exhibit peritectic reaction during solidification. In order to investigate the solidification behavior of Zn-rich Zn–Cu peritectic alloy containing nominally up to 7.8 wt.% Cu, a series of laser surface remelting experiments were performed. With the increase in growth velocity, Zn–Cu alloys with Cu content below 3.0wt.% showed an evolutional sequence from low-velocity η planar interface?→?lamellar structures?→?η shallow cells and finally to high-velocity η planar interface. The Zn-4.0 wt.%Cu alloy showed a similar transitional sequence except that irregular η cells appeared when low-velocity planar interface became unstable. In contrast, ε cell/dendrite was the typical microstructure of the Zn-7.8 wt.% Cu alloy within the whole scanning velocity range. Based on the maximum interface temperature criterion, a eutectic growth model under rapid solidification conditions (TMK model) and a self-consistent numerical model for the cellular and dendrite growth were applied to establish a phase and microstructure pattern selection map, which drew a clear whole picture of the relationship between phase/microstructure and solidification conditions of this series of alloys. Regarding the microstructure feature, our investigation revealed the range of the solidification velocity and chemical composition of lamellar structures as dominant microstructure and their lamellar spacing displayed a considerable range of the average value as a function of growth velocity. The relationship between the lamellar spacing and the growth velocity was further analyzed by using the TMK eutectic model, and the results showed the same overall trend as the experimental results.

Graphical abstract

A phase and microstructure pattern selection map of Zn-rich Zn–Cu peritectic alloys. Regression analysis of the average spacing of lamellar structures.

  相似文献   

2.
This study investigated the effects of cooling rate during solidification, heat treatment, and the addition of Mn and Sr on the formation of intermetallic phases in Al–11Si–2.5Cu–Mg alloys. Microstructures were monitored using optical microscopy and EPMA techniques. The results reveal that the volume fractions of intermetallic phases are generally much lower in the furnace-cooled samples than in the air-cooled ones due to the dissolution of the β-AlFeSi and Al2Cu phases during slow cooling at critical dissolution temperatures. Strontium additions increased the volume fraction of the Al2Cu phase in the as-cast conditions at low and high cooling rates, as well as at varying ranges of Mn levels. Platelets of the β-AlFeSi phase were to be observed in the microstructure of the as-cast air-cooled samples with a DAS of 40 μm at both Mn levels, while none of these particles were to be found in the furnace-cooled samples with a DAS of 120 μm. Sludge particles were observed in almost all of the air-cooled alloys with sludge factors of between 1.4 and 1.9. These particles, however, were not observed in the furnace-cooled alloys with similar sludge factors. Solution heat treatment coarsens the Si particles in the non-modified alloys under both sets of cooling conditions studied. In the Sr-modified alloys, solution treatment has varied effects depending on the cooling rate and the level of Mn present.  相似文献   

3.
In order to examine experimentally the growth behavior of Nb3Sn during reactive diffusion between Nb and a bronze with the α + β two-phase microstructure, a sandwich (Cu–Sn–Ti)/Nb/(Cu–Sn–Ti) diffusion couple was prepared from pure Nb and a ternary Cu–Sn–Ti alloy with concentrations of 9.3 at.% Sn and 0.3 at.% Ti by a diffusion bonding technique. Here, α is the primary solid-solution phase of Cu with the face-centered cubic structure, and β is the intermediate phase with the body-centered cubic structure. The diffusion couple was isothermally annealed at temperatures between T = 923 and 1,053 K for various times up to 843 h. Owing to annealing, the Nb3Sn layer is formed along each (Cu–Sn–Ti)/Nb interface in the diffusion couple, and grows mainly into Nb. Hence, the migration of the Nb3Sn/Nb interface governs the growth of the Nb3Sn layer. The mean thickness of the Nb3Sn layer is proportional to a power function of the annealing time. The exponent of the power function is close to unity at T = 923 K, but takes values of 0.8–0.7 at T = 973–1,053 K. Consequently, the interface reaction at the migrating Nb3Sn/Nb interface is the rate-controlling process for the growth of the Nb3Sn layer at T = 923 K, and the interdiffusion across the Nb3Sn layer as well as the interface reaction contributes to the rate-controlling process at T = 973–1,053 K. Except the effect of Ti, the growth rate of the Nb3Sn layer is predominantly determined by the activity of Sn in the bronze and thus the concentration of Sn in the α phase. As a result, the growth rate is hardly affected by the volume fraction of the β phase, though the final amount of the Nb3Sn layer may depend on the volume fraction.  相似文献   

4.
The martensitic transformation and shape memory effect (SME) in a β Cu–Zn alloy containing various amounts of α1 plates have been investigated. The results showed that the characteristic temperature of martensitic transformation decreased as the isothermal aging time increased. The crystal structure, twin relationships between martensite variants, and characteristics of martensitic transformation of the β Cu–Zn alloy were not affected by the existence of α1 plates. However, the α1 plates were distributed homogeneously in the parent phase and functioned as grain boundaries, hindering the progress of martensite variants, and reducing the effective grain size of the parent phase and the size of self-accommodating plate groups formed upon cooling. In addition, the strain recovery due to the SME decreased as the isothermal aging time increased (the quantity of α1 plate increased) and/or imposed prestrain increased. Nevertheless, the SME mechanism in the β Cu–Zn alloy containing α1 plates was not affected by the presence of the α1 plates.  相似文献   

5.
In this study, the influence of magnesium content on thermal and structural parameters during the unsteady-state unidirectional solidification of Al–Mg alloys is analyzed. Using a special device, Al–Mg alloys containing 5, 10, and 15 wt% Mg were submitted to unidirectional solidification. Using a data acquisition system, the temperature variations along the casting during solidification were measured. From these results, the variations of solidification parameters as growth rate of dendrite tips, thermal gradient, cooling rate, and local solidification time were determined. The variation of global heat transfer coefficient at metal/mould interface was estimated through the adjustment of experimental temperature variation close to the interface and numerical predictions. Primary and secondary dendrite arms spacing variations during solidification were measured by optical microscopy. From these results, comparative analysis were developed to determine the influence of magnesium content.  相似文献   

6.
Brazing of a nitrogen-containing duplex stainless steel was preformed using a nickel-based filler metal (Ni-4.5wt.%, Si-3.2wt.%, B). The microstructure of the brazed joint was investigated using scanning electron microscopy, energy dispersive spectrometry, electron probe microanalyzer, and layer-by-layer X-ray diffraction analysis. The results indicated that before completion of isothermal solidification, BN, Ni3B and Ni3Si precipitates formed at the interface, in the athermally solidified zone and isothermally solidified zone, respectively. After isothermal solidification, only γ-Ni phase appeared in the brazed interlayer. The appearance of hardness peak values in the athermally solidified zone and the interface most probably corresponded to the formation of Ni3B and BN, respectively.  相似文献   

7.
Transient liquid-phase bonding of a duplex stainless steel was performed with a Ni–Cr–B insert alloy. The microstructure of the joint region was investigated by cross-sectional and layer-by-layer characterization. According to the experimental studies, prior to completion of isothermal solidification, the bond microstructure can be expressed as γ-Fe + δ-Fe/γ-Fe + δ-Fe + BN/γ-Ni(Fe) + BN/γ-Ni + Cr-rich borides/γ-Ni + Ni3B + Cr-rich borides (CrB, CrB2, Cr2B3, Cr3B4, Cr5B3 and CrB4), from the base metal side to the bonded-interlayer side. Complete isothermal solidification occurred at 1090 °C within 3600 s. Only the γ-Ni solid solution phase was present in the bonded interlayer, and BN precipitates were not removed after isothermal solidification. The formation of secondary-phase precipitates might be responsible for the presence of peak microindentation hardness in the bond region.  相似文献   

8.
The directional solidification of Pb-Sn alloys   总被引:2,自引:0,他引:2  
Directional solidification experiments have been carried out on different Pb-Sn alloys as a function of temperature gradient G, growth rate V and cooling rate GV. The specimens were solidified under steady state condition with a constant temperature gradient (50 °C/cm) at a wide range of growth rates ((10–400) × 10–4 cm/s) and with a constant growth rate (17 × 10–4 cm/s) at a wide range of temperature gradient (10–55 °C/cm). The primary dendrite arm spacing, 1, and secondary dendrite arm spacing, 2, were evaluated. This structure parameters were expressed as functions of G, V and GV by using the linear regression analysis. The results were in good agreement with the previous works.  相似文献   

9.
Abstract

Microstructures of the two ternary eutectic alloys of the Bi–Cd–In system were studied using slow unidirectional solidification, followed by quenching to form a representative solid/liquid interface for subsequent observation. The eutectic reactions were found to take the form L?BiIn+BiIn2+Cd at 77.5°C and L?BiIn2+?+Cd at 61.5°C. The 77.5°C eutectic was observed to be of the faceted (BiIn)–faceted (Cd)–non-faceted (BiIn2) type, while all three phases of the 61.5°C eutectic showed faceting. The BiIn and BiIn2 phases of the 77.5°C eutectic formed a quasiregular microstructure with the Cd phase growing relatively independently. The phases of the 61.5°C eutectic tended to form a lamellar microstructure with a BiIn2?–Cd–?–BiIn2 phase sequence. Both eutectics were observed to obey the usual phase spacing law, λ2R=constant, where λ is the phase spacing and R is the growth rate.  相似文献   

10.
Solidification-related phenomena and the properties of the final product are strongly influenced by the developing dendritic microstructure, which is defined e.g. by the secondary dendrite arm spacing. In the past, different experimental set-ups were applied and subsequently the secondary dendrite arm spacing of certain steel grades was measured. However, it is difficult to compare the proposed relations based on either the local solidification time or the cooling rate, and they also vary over a wide range. Therefore, the present study systematically investigates the effect of carbon on the secondary dendrite arm spacing using in situ solidification experiments with accurately defined solidification conditions. The parameter K in the empirical equation was determined as a function of carbon, using an iterative procedure to calculate the local solidification time and the measured secondary dendrite arm spacings. Furthermore, these results were discussed and compared with theoretical models from the literature.  相似文献   

11.
The microstructure and solidification behavior of Cu–Ni–Si alloys with four different Cu contents was studied systematically under near-equilibrium solidification conditions. The microstructures of these Cu–Ni–Si alloys were characterized by SEM and the phase composition was identified by XRD analysis. The phase transition during the solidification process was studied by DTA under an Ar atmosphere. The results show that the microstructure and solidification behavior is closely related to the composition of Cu–Ni–Si alloys. The microstructure of Cu–Ni–Si alloys with higher than 40% Cu content consists of primary phase α-Cu(Ni, Si) and eutectic phase (β1-Ni3Si + α-Cu(Ni,Si).When the Cu content is about 40%, only the eutectic phase (β1-Ni3Si + α-Cu(Ni,Si)) is present. DTA analysis shows there are three phase transitions during every cooling cycle of alloys with higher than 40% Cu content, but only one for 40% Cu content. Cu–Ni–Si alloy with 40% Cu solidifies by a eutectic reaction, but Cu–Ni–Si alloys with higher than 40% Cu content solidify as a hypoeutectic reaction.  相似文献   

12.
Abstract

An earlier procedure for calculating homogenization during the solidification of binary alloys is extended to include more realistic assumptions for dendrite arm coarsening kinetics and the partition coefficient. Dendrite coarsening is shown to follow the usual exponential function which differs from that relating final dendrite arm spacing λf to local solidification time. This function is not sensitive to the assumption for the partition coefficient or the choice of initial spacing (between λ0 = 0 and λf/2). The homogenization parameter H (normalized difference between the minimum possible concentration and the actual minimum concentration in the centre of the dendrite arm) changes only slightly with the partition coefficient. With the average coarsening parameter used by earlier authors, homogenization is overestimated and a lower amount of the non-equilibrium phase is obtained. The experimental data available for Al–Cu alloys in the range 2–20 wt-% Cu are very closely predicted by the numerical procedure used here.

MST/435  相似文献   

13.
The self-propagating high-temperature synthesis (SHS) reactions can take place in Cu–Ti–Si systems with Cu additions of 10–50 wt.%, and the products only consist of Ti5Si3 and Cu phases, without any transient phase. In Ti–Si system, most of the Ti5Si3 grains synthesized exhibit the polygon-shaped coarse appearance with an obviously sintered morphology. When Cu content increases from 10 to 50 wt.%, however, the Ti5Si3 exhibits cobblestone-like shape with a relatively smooth surface, and its average size decreases significantly from 15 to 2 μm or less. The formation mechanism of Ti5Si3 in Cu–Ti–Si system is characterized by the solution, reaction and precipitation processes. Furthermore, the addition of Cu has a great influence on the volume change between green and reacted preforms. The volume change increases with Cu content increasing from 0 to 20 wt.%, and then decreases with the content further increasing from 20 to 50 wt.%. The addition of Cu to Ti–Si system significantly decreases the onset temperature of the reaction during differential scanning calorimetry process, which is even much lower than the α → β transition temperature of Ti (882 °C), suggesting that the reaction could be greatly facilitated by Cu addition. As a result, the role of Cu serves not only as a diluent but also as a reactant and participates in the self-propagating high-temperature synthesis reaction process.  相似文献   

14.
15.
Non modified and Ag-modified eutectic Sn-0.7Cu solder alloys were directionally solidified under transient heat flow conditions. The microstructure of the Sn-0.7Cu alloy has been characterized and the present experimental results include the cell/primary dendrite arm spacing (λ1) and its correlation with: the tip cooling rate ( $\mathop T\limits^{ \bullet }$ ) during solidification, ultimate tensile strength (σu) and elongation to fracture (δ). Distinct morphologies of intermetallic compounds have been associated with the solidification cooling rate for both alloys examined. For the Sn-0.7Cu alloy, cellular regions were observed to occur for cooling rates lower than 0.9 K/s, being characterized by aligned eutectic colonies. On the other hand, the alloy containing 2.0 wt %Ag enabled the launch of tertiary branches within the dendritic arrangement. The comparison of results allows stating that finer solder microstructures are shown to be associated with higher ultimate tensile strengths (σu) for both alloys although a more complex microstructure was found for the SAC alloy. In contrast the elongation (δ) exhibited opposite tendencies. The growth of coarse Ag3Sn fibers and platelets within interdendritic regions seems to contribute for the reduction on ductility observed for the SAC alloy.  相似文献   

16.
After establishing steady-state directional growth, aluminum-6 wt.% silicon alloys were subjected to a series of programmed accelerations and decelerations. The response of the primary dendrite arm spacing (λ1), the eutectic spacing (λE), and the primary dendrite trunk diameter (d) to the imposed rate changes were evaluated and compared with results from constant growth velocity experiments. It was found that both λE and d responded to a continually changing growth rate, whereas λ1 did not. The implication of using these microstructural features to characterize solidification histories is discussed.  相似文献   

17.
In this study, Si3N4 ceramic was jointed by a brazing technique with a Cu–Zn–Ti filler alloy. The interfacial microstructure between Si3N4 ceramic and filler alloy in the Si3N4/Si3N4 joint was observed and analyzed by using electron-probe microanalysis, X-ray diffraction and transmission electron microscopy. The results indicate that there are two reaction layers at the ceramic/filler interface in the joint, which was obtained by brazing at a temperature and holding time of 1223 K and 15 min, respectively. The layer nearby the Si3N4 ceramic is a TiN layer with an average grain size of 100 nm, and the layer nearby the filler alloy is a Ti5Si3Nx layer with an average grain size of 1–2 μm. Thickness of the TiN and Ti5Si3Nx layers is about 1 μm and 10 μm, respectively. The formation mechanism of the reaction layers was discussed. A model showing the microstructure from Si3N4 ceramic to filler alloy in the Si3N4/Si3N4 joint was provided as: Si3N4 ceramic/TiN reaction layer/Ti5Si3Nx reaction layer/Cu–Zn solution.  相似文献   

18.
Microstructural characterization of α1-plate and γ2 phase precipitated in hypoeutectoid Cu–10 wt.%Al–0.8 wt.%Be shape-memory alloy (SMA) aged at 200 °C for different periods of time (20–160 h) is researched in this study. High-resolution transmission electron microscope (HRTEM) was employed to investigate the α1-plate with 18R long period stacking order structure (LPSO) in the SMA aged for 20 h. According to the atomic shuffling revealed in HRTEM-micrograph, the atomic model of the 18R LPSO is proposed. The quantitative mapping of electron energy loss spectrometry shows that the α1-plates in the SMA aged for 160 h contain lower aluminum concentration than the parent phase matrix. The lattice image of the nanometer-sized γ2 phase precipitated homogeneously in the SMA aged for 160 h is also revealed by using HRTEM. Precipitation of the nanometer-sized γ2 phase cannot be impeded by means of the addition of beryllium and quenching, and such precipitate does not grow up in the SMA aged for periods of time less than 160 h.  相似文献   

19.
The theoretical strength and structural response of FCC crystal Cu under uniaxial loading along one edge of a unit cell have been investigated with MAEAM. The stability criteria of the tetragonal system are extended to the generally stressed orthogonal system. The results show that, even if an orthorhombic path is applied, the deformation is spontaneously along the tetragonal Bain path till tensile elastic limit at stretch λ1 = 1.085 where the Born criterion is violated. The branched orthogonal path is preferred over the conventional tetragonal Bain path from minimization of the stress σ1 or the energy E. Although a stress-free BCC phase with the local maximum energy of −3.460 eV appearing either in compressive region or in tensile region is unstable and would slip spontaneously into a near neighbor stress-free mBCT phase with the local minimum energy of −3.461 eV, the initial FCC phase with the lowest energy of −3.489 eV is the most stable in correspondence with the actual behavior of pure Cu. Furthermore, the calculated elastic moduli of FCC phase are in good agreement with the experimental values. The stable region ranges from −2.89 GPa to 7.252 GPa in the theoretical strength or from 0.912 to 1.085 in the stretch λ1 correspondingly.  相似文献   

20.
Effects of trace amount of rare earth element Pr on properties and microstructure of Sn–0.7Cu–0.05Ni solder were investigated in this paper. The solderability of Sn–Cu–Ni–xPr alloy and shear strengh of Sn–Cu–Ni–xPr soldered micro-joints were determined by means of the wetting balance method and shear test, respectively. Moreover, microstructure of solder alloys bearing Pr, as well as intermetallic compound (IMC) layer formed at solder/Cu interface after soldering were observed. It was concluded that the major benefits of rare earth element Pr on Sn–Cu–Ni lead-free solder are: improving solderability, refining microstructure, and depressing IMC (IMC) growth, which exhibited improved mechanical properties. It also revealed that (Cu,Ni)6Sn5 is the majority IMC phase at the interface of Sn–Cu–Ni–xPr/Cu solder joints. Ni added into the solder effectively suppressed the growth of Cu3Sn and consequently also the total IMC layer thickness. Above all, the thickness and morphology of the interfacial (Cu,Ni)6Sn5 IMC were optimized due to alloying Pr. It can be inferred that Pr and Ni would play an important role in improving the reliability of Sn–Cu–Ni lead-free solder joints.  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