Journal of Materials Science & Technology  2020 , 44 (0): 201-208 https://doi.org/10.1016/j.jmst.2019.10.038

Research Article

Microstructure formation and electrical resistivity behavior of rapidly solidified Cu-Fe-Zr immiscible alloys

Xiaojun Sunab, Jie Heab*, Bin Chenab, Lili Zhanga, Hongxiang Jianga, Jiuzhou Zhaoab, Hongri Haoa

a Institute of Metal Research, Chinese Academy of Sciences, Shenyang 110016, China
b School of Materials Science and Engineering, University of Science and Technology of China, Shenyang 110016, China

Corresponding authors:   * E-mail address: jiehe@imr.ac.cn (J. He).

Received: 2019-09-19

Revised:  2019-10-9

Accepted:  2019-10-12

Online:  2020-05-01

Copyright:  2020 Editorial board of Journal of Materials Science & Technology Copyright reserved, Editorial board of Journal of Materials Science & Technology

More

Abstract

The immiscible Cu-Fe alloy was characterized by a metastable miscibility gap. With the addition element Zr, the miscibility gap can be extended into the Cu-Fe-Zr ternary system. The effect of the atomic ratio of Cu to Fe and Zr content on the behavior of liquid-liquid phase separation was studied. The results show that liquid-liquid phase separation into Cu-rich and Fe-rich liquids took place in the as-quenched Cu-Fe-Zr alloy. A glassy structure with nanoscale phase separation was obtained in the as-quenched (Cu0.5Fe0.5)40Zr60 alloy sample, exhibiting a homogeneous distribution of glassy Cu-rich nanoparticles in glassy Fe-rich matrix. The microstructural evolution and the competitive mechanism of phase formation in the rapidly solidified Cu-Fe-Zr system were discussed in detail. Moreover, the electrical property of the as-quenched Cu-Fe-Zr alloy samples was examined. It displays an abnormal change of electrical resistivity upon temperature in the nanoscale-phase-separation metallic glass. The crystallization behavior of such metallic glass has been discussed.

Keywords: Immiscible alloys ; Liquid-liquid phase separation ; Rapid solidification ; Microstructure ; Electrical resistivity behavior

0

PDF (3557KB) Metadata Metrics Related articles

Cite this article Export EndNote Ris Bibtex

Xiaojun Sun, Jie He, Bin Chen, Lili Zhang, Hongxiang Jiang, Jiuzhou Zhao, Hongri Hao. Microstructure formation and electrical resistivity behavior of rapidly solidified Cu-Fe-Zr immiscible alloys[J]. Journal of Materials Science & Technology, 2020, 44(0): 201-208 https://doi.org/10.1016/j.jmst.2019.10.038

1. Introduction

Alloys with a metastable miscibility gap in the range of the undercooled liquid state have been of increasing interest [[1], [2], [3], [4], [5], [6]]. Many of them are excellent candidates for industrial applications. The Cu-Fe alloy, as a kind of high strength and high electric conductivity materials [[7], [8], [9], [10], [11]], is a typical peritectic system. From the Cu-Fe binary alloy diagram [12,13], it has a nearly flat liquidus line and a retrograde solidus line and hence displays a thermodynamic tendency toward the formation of a miscibility gap in undercooled liquid state. The effect of the cooling rate, undercooling degree and gravity level on the microstructure evolution of the Cu-Fe system was studied through experiments such as gas atomization [[14], [15], [16]], containerless solidification [17,18] and drop tube [19]. Novel structure with self-organized core/shell was detected in these powders, which are promising solder balls for modern electronic packaging technology. Much attention has been paid to the microstructure formation to clarify the kinetics of metastable phase separation in the undercooled binary liquid Cu-Fe alloys [[20], [21], [22], [23], [24], [25]]. It was indicated that the formation of such structure are attributed to the droplets spacial migration and coagulation and the wetting behavior [[14], [15], [16],26,27]. More recently, Li and co-worker [22,23] analyzed the mechanism of the secondary liquid-liquid phase separation (LLPS) in the Cu80Fe20 alloy subjected different cooling conditions, and examined soft ferromagnetic characteristics of the Cu80Fe20 alloy.

The metastable miscibility gap of the binary Cu-Fe system may be extended into some ternary and multicomponent systems [28,29]. For some ternary Cu-Fe-X (X = Pb, Si, C, etc.) systems, the third addition can enhance the LLPS [30,31]. The liquid-liquid hierarchical separation phenomenon was detected in the ternary Cu-Fe-Pb alloy, having new potential applications in the separation and recovery of the secondary complex metal resource of printed circuit boards [32,33]. In contrast, a metastable LLPS into Fe-rich and Cu-rich liquids was examined for the Cu-Fe-Z (Z = Ni, Co, Cr, Mn, etc.) alloy [34]. The phenomenon of phase separation was utilized to enhance their mechanical properties of some multicomponent systems (i.e., high-entropy alloys [35]). The elements Cu, Fe and Zr are the basic constituents for some engineering materials [[36], [37], [38], [39]] such as high-strength and high-conductivity copper alloys, glass-forming alloy systems, etc. As known, enthalpies of mixing (ΔHmix) of Cu-Fe, Cu-Zr, and Fe-Zr are +12 kJ mol-1, -23 kJ mol-1 and -25 kJ mol-1, respectively [40]. The negative and positive enthalpies mean strong attractive and repulsive interactions between atoms, and may lead to changes of local structure [[41], [42], [43], [44]], which influence the properties of such materials. In this work, the microstructure formation of the rapidly solidified Cu-Fe-Zr ternary alloys was studied. The effect of Zr content, atomic ratio of Cu to Fe, and cooling rate on the structure of the Cu-Fe-Zr ternary system has been clarified. What’s more, an abnormal change of electrical resistivity upon temperature in the as-quenched Cu-Fe-Zr alloys was detected for the first time. The crystallization behavior has been discussed in detail. It is of significance in understanding the property of complex multicomponent Cu-Fe-Zr-based metal materials.

2. Material and methods

The ingots of master alloys were prepared from Cu, Fe and Zr metals with purity better than 99.95% by arc melting under a Ti-gettered purified argon atmosphere in water-cooled copper crucible. The alloy ingots were melted at least four times to ensure homogeneous. The 5-10 g master alloy was remelted in a quartz-glass tube with a 0.7 mm orifice and ejected through a nozzle onto the copper wheel with an overpressure of 50 kPa. Rapidly quenched ribbon was fabricated by single-roller melting-spinning method with the surface velocity of the copper roller is about 27 m s-1. The cylindrical specimens with 2 mm in diameter were prepared by remelting the master alloy in quartz-glass tubes and subsequent ejecting of the melt with an overpressure of 50 kPa through a nozzle into copper mold.

X-ray diffraction (XRD) analysis was performed for the samples by using CuKa radiation for a 2θ range of 20°-80°. The microstructure of the rapidly solidified samples was characterized by scanning electron microscopy (SEM), transmission electron microscopy (TEM) linked with an STEM detector. At heating and cooling rates of 20 K min-1, the thermal properties of the samples were analyzed by differential scanning calorimeter (DSC) under argon atmosphere. The change of the electrical resistivity upon temperature for the sample with dimension of about 1.5 mm × 1.0 mm × 0.2 mm was measured using a standard four-probe technique with a heating rate of 10 K min-1.

3. Results and discussion

3.1. Correlations between microstructure and composition

3.1.1. Effect of atomic ratio of Cu to Fe

Fig. 1 shows the microstructures of the as-quenched (Cu1-xFex)90Zr10 (x = 0.45, 0.5, 0.6) alloys. Obviously, the microstructure is correlated to the atomic ratio of Cu to Fe. When the Fe content x in (Cu1-xFex)90Zr10 alloys were 0.45 and 0.5, the two-phase interconnected structure was observed, as shown by the bright and dark regions in Fig. 1(a) and (b). The EDS analysis reveals that the dark and bright regions are Fe-rich and Cu-rich phases, respectively. Average compositions Cu32.13Fe56.98Zr10.89 and Cu68.48Fe23.53Zr7.98 for the two regions in as-quenched (Cu0.5Fe0.5)90Zr10 alloys were detected, respectively, indicating that the solute element Zr dissolves in both Cu-rich and Fe-rich regions. From their microstructural characteristics, it can be concluded that the LLPS into Cu-rich and Fe-rich liquids occurs during the rapid cooling. In contrast, a particle-isolated structure forms in the as-quenched (Cu0.4Fe0.6)90Zr10 alloy sample, as shown in Fig. 1(c). The characteristic size and volume fraction estimated by standard stereological methods of Cu-rich particles are 0.2‒0.8 μm and 17.58% ± 0.2% (Fig. 1(d)). The miscibility gap of the binary Cu-Fe alloy can be extended to the ternary Cu-Fe-Zr alloy system. The LLPS can occur by two mechanisms, i.e., by liquid-state spinodal decomposition or by liquid-state nucleation and growth. The liquid-state spinodal decomposition normally results into the interconnected structure shown in Fig. 1(a) and (b). The particle-isolated microstructure in Fig. 1(c) is attributed to the LLPS by nucleation and growth. Comparing Fig. 1(a), (b) with Fig. 1(c) suggests that the microstructure of the as-quenched (Cu1-xFex)90Zr10 alloys is tunable by changing the atomic ratio of Cu to Fe.

Fig. 1.   SEM images of the as-quenched (a) (Cu0.55Fe0.45)90Zr10, (b) (Cu0.5Fe0.5)90Zr10, and (c) (Cu0.4Fe0.6)90Zr10 alloys, (d) size distribution of the particles in the as-quenched (Cu0.4Fe0.6)90Zr10 alloy.

3.1.2. Effect of Zr content

Fig. 2 shows the XRD patterns and DSC curves of the as-quenched (Cu0.5Fe0.5)100-xZrx (x = 10, 20, 40, 60, 80) alloys. For x = 10, the alloy sample has a completely crystalline structure and is mainly composed of Cu and Fe2Zr phase. With the further increase of Zr content, an obvious broad diffraction peak appears in the XRD patterns of as-quenched (Cu0.5Fe0.5)60Zr40 alloys. Sharp peaks of crystalline phases superimpose on the broad scattering hump characteristic of an amorphous phase. When the Zr content x reaches 60, the XRD pattern only exhibits a broad diffraction maximum without any crystalline reflection, indicating that the structure of the as-quenched (Cu0.5Fe0.5)40Zr60 alloy sample is fully amorphous. It indicates that an increase of Zr content can obviously improve the glass-forming ability (GFA) of (Cu0.5Fe0.5)100-xZrx system. The DSC scans of the as-quenched (Cu0.5Fe0.5)100-xZrx alloys are shown in Fig. 2(b). During the melting stage, it is clear that there are two endothermic reactions, i.e., the melting of Cu-rich phase and that of Fe-rich phase, as marked by the dotted frames in Fig. 2(b). The DSC curves of the as-quenched (Cu0.5Fe0.5)100-xZrx (x = 40, 60, 80) alloys exhibit obvious crystallization exothermic peak. It further verifies the existence of amorphous phase in these samples. The crystalline peaks have a left shift with the increase of the Zr content, which means a decrease of thermal stability.

Fig. 2.   (a) XRD patterns and (b) DSC heating scans of as-quenched (Cu0.5Fe0.5)100-xZrx alloys.

Fig. 3(a) reveals the STEM image of the as-quenched (Cu0.5Fe0.5)40Zr60 alloys. The (Cu0.5Fe0.5)40Zr60 alloys exhibits a heterogeneous structure with the white Cu-rich nanoparticles embedded in the gray Fe-rich matrix. The characteristic size of the Cu-rich nanoparticle is 2-8 nm, approximating to a log-normal distribution (Fig. 3(b)). The population density and volume fraction of nanoparticles are estimated by standard stereological methods to be 5.37 × 1024 m-3 and 10.4% ± 0.2%. The selected-area diffraction pattern (SAED) shown by an inset in Fig. 3(a) indicates that the structure of the as-quenched (Cu0.5Fe0.5)40Zr60 alloy is fully glassy. The HRTEM images indicate no evidence for nanocrystals (NCs) in any of the samples and it also shows that long strips of microstructure exists in addition to particle-isolated microstructure (see Fig. 3(c) and (d)).

Fig. 3.   (a) STEM image of the as-quenched (Cu0.5Fe0.5)40Zr60 alloy (the inset is the SAED pattern), (b) size distribution of the glassy nanoparticles in the as-quenched (Cu0.5Fe0.5)40Zr60 alloy, (c) and (d) HRTEM images of the typical nanoparticles in the as-quenched (Cu0.5Fe0.5)40Zr60 alloy, exhibiting different appearances.

Considering that the (Cu1-xFex)40Zr60 alloy system has a better GFA, we further studied the effect of the atomic ratio of Cu to Fe on the structure of the as-quenched (Cu1-xFex)40Zr60 alloys. Fig. 4 shows the XRD patterns of the as-quenched (Cu1-xFex)40Zr60 (x = 0.3, 0.5, 0.6) alloys. Obviously, the GFA of the (Cu1-xFex)40Zr60 alloy is preferable when atomic ratio of Cu to Fe is between 1:1 and 7:3. For (Cu0.4Fe0.6)40Zr60 alloys, sharp peaks of crystalline Fe2Zr phases superimpose on the broad scattering hump characteristic of an amorphous phase. The position of broad peak in single-phase Zr-Cu and Zr-Fe amorphous alloys is almost identical, and therefore, there is only one broad peak in XRD patterns of Cu-Fe-Zr system, as shown in Fig. 4.

Fig. 4.   XRD patterns of as-quenched (Cu1-xFex)40Zr60, Zr60Cu40, and Zr75Fe25 alloys.

The TEM image of the as-quenched (Cu0.7Fe0.3)40Zr60 alloy is shown in Fig. 5. The as-quenched (Cu0.7Fe0.3)40Zr60 alloy exhibits an structure with the white Fe-rich spheres embedded in the gray Cu-rich matrix. The phase separation into the Cu-rich and Fe-rich liquids took place during the rapid cooling. The SAED presents halo diffraction intensity with a few diffraction spots, indicating the as-quenched (Cu0.7Fe0.3)40Zr60 alloy contains NCs. The examination indicates that the NCs are orignated from the black nanospheres shown in the Fig. 5. The diameter of the white Fe-rich spheres is 20-50 nm. Comparing the structure of the (Cu0.7Fe0.3)40Zr60 alloy with that of (Cu0.5Fe0.5)40Zr60 alloy, we can conclude that the scale of phase separation and the GFA of the (Cu1-xFex)40Zr60 system could be tunable by changing the atomic ratio of Cu to Fe.

Fig. 5.   TEM image of as-quenched (Cu0.7Fe0.3)40Zr60 alloys.

3.2. Microstructure formation

From the viewpoint of thermodynamics, the liquid-solid phase transformation (LSPT) is possible if the difference Gibbs free energy (ΔGSL) between the final solid state and initial liquid state is negative:

$\Delta G_{SL}=G_{S}-G_{L}<0$ (1)

where GS and GL are Gibbs free energy of the solid solution and liquid and can be described by:

$G_{S}=\sum_{i} 0_{G}x_{i}+RT \sum_{i} x_{i} ln x_{i}+E_{G}+\Delta G_{mag}$ (2)

$G_{L}=\sum_{i} 0_{G}x_{i}+RT \sum_{i} x_{i} ln x_{i}+E_{G}$ (3)

where R is the gas constant, T is the absolute temperature (K), xi is the mole fraction of component i, ΔGmag is the magnetic contribution to the free energy and EGm is the excess free energy. Also, the difference in the Gibbs free energy between homogeneous liquid and the completely separated liquid can be considered as a driving force for LLPS. The magnitude of driving force determines the probability of phase transformation. Fig. 6 reveals the Gibbs free energy of the undercooled Cu-Fe melt at 1500 K. Although the Gibbs free energy of γ-Fe is lower than that of the melt when Fe content is higher than 41% (Fig. 6(a)), the Gibbs free energy change on primary solidification and LLPS (Fig. 6(b)) explicitly shows that LLPS occurs preferentially in the melt before its solidification when Fe content is lower than 54%. Otherwise, the γ-Fe phase resulting from a liquid-solid transition may primarily form.

Fig. 6.   Gibbs free energies of the undercooled Cu-Fe melt at T = 1500 K: (a) Gibbs free energies GL and Gγ-Fe of the melt and γ-Fe, (b) Gibbs free energy difference for phase separation in liquid and solidification of γ-Fe.

The miscibility gap of the (Cu1-xFex)90Zr10 system was calculated using thermodynamic data [13,[45], [46], [47]]. The liquid miscibility gap exhibits a pronounced asymmetry (Fig. 7(a)), whose critical point temperature (Tc) and composition are approximately 1379 K and Cu54Fe36Zr10, respectively. The liquidus temperatures for the (Cu1-xFex)90Zr10 (x = 0, 0.2, 0.3, 0.4, 0.5, 0.6, 0.7, 0.8, and 1) alloys were determined by DSC, and the results are shown in Fig. 7. Similar to the binary Cu-Fe system, there exists the competitive behavior of phase formation in the as-quenched (Cu1-xFex)90Zr10 alloys. The compositions (Cu0.55Fe0.45)90Zr10 and (Cu0.5Fe0.5)90Zr10 are close to the critical point composition Cu54Fe36Zr10. Combining DSC analyses with thermodynamic calculations shows the undercooling degree required for LLPS is about 110 K for the Cu54Fe36Zr10 alloy. During rapid quenching, the spinodal decomposition is more possible for these alloys, leading to two-phase interconnected structure (Fig. 1(a) and (b)). The composition (Cu0.4Fe0.6)90Zr10 is far away from the critical point composition, exhibiting a larger undercooling degree required for the spinodal decomposition. The LLPS may occur by the liquid-state nucleation and growth, and thus result in the particle-isolated microstructure (Fig. 1(c)). If the melt was not undercooled into the liquid miscibility gap of the (Cu1-xFex)90Zr10 alloys, the LSPT may primarily occur during cooling. Fig. 7(b) and (c) shows the microstructure of the (Cu0.4Fe0.6)90Zr10 and (Cu0.8Fe0.2)90Zr10 alloy samples after the DSC circle at a cooling rate of 20 K s-1. The Fe dendrites primarily precipitate at about 1525 K and then Fe2Zr phase forms at about 1485 K during the (Cu0.4Fe0.6)90Zr10 alloy cooling, as shown in Fig. 7(a). The residual liquid was enriched by Cu solidifies at about 1328 K. For the (Cu0.7Fe0.3)90Zr10 and (Cu0.8Fe0.2)90Zr10 alloys, besides Fe2Zr and Cu phases, the formation of Cu5Zr phase corresponding to the exothermic reaction at 1250 K was detected, as shown Fig. 7(a) and (c).

Fig. 7.   (a) Calculated metastable miscibility gap and DSC curves of the (Cu1-xFex)90Zr10 system, (b) and (c) SEM images of the (Cu0.4Fe0.6)90Zr10 and (Cu0.8Fe0.2)90Zr10 alloy samples after a cooling DSC circle.

As mentioned above, the microstructure of the as-quenched (Cu1-xFex)40Zr60 alloys may be composed of glassy matrix, glassy particles, and crystalline precipitates. It means that the competition between LLPS, LSPT and liquid-glass transition (LGT) plays a decisive role in the final phase formation. Here, the competitive mechanism of phase formation in the as-quenched (Cu1-xFex)40Zr60 alloys is analyzed. Fig. 8(a) shows the miscibility gap and the liquidus temperature of the (Cu1-xFex)40Zr60 alloys. For as-quenched (Cu0.5Fe0.5)40Zr60 alloys, LLPS takes place when alloy melt is undercooled into the miscibility gap during rapid quenching and then LGT occurs at Tg = 615 K, leading to the formation of a microstructure with glassy matrix embedded by glassy particles (Fig. 3(a)).

Fig. 8.   (a) Calculated metastable miscibility gap of the (Cu1-xFex)40Zr60 system, (b) schematic diagram of viscosity changing with temperature. TL and Ts present the liquidus temperature and onset temperature of LLPS, respectively.

For the (Cu0.7Fe0.3)40Zr60 alloy, the LLPS into Cu-rich and Fe-rich liquids takes place during the rapid quenching. The driving force for LSPT is higher than that for LGT because the Cu content of the Fe-rich droplets formed by LLPS was obviously lower than that of Cu-rich matrix in the periphery of Fe-rich spheres. As shown in the rectangular area in Fig. 5, the black Cu-rich NCs generated from LSPT prefer to be conjoined with of the white Fe-rich glassy sphere, which indicates that the interface between the Fe-rich droplets and Cu-rich liquid matrix may act as nucleation sites for the Cu-rich NCs. This phenomenon also were detected in the rapidly solidified binary Cu-Co and Cu-Fe alloys [48,49]. Consequently, at this moment, the as-quenched (Cu0.7Fe0.3)40Zr60 system consists of the Cu-rich liquid matrix, Fe-rich droplets, and Cu-rich NCs. The Cu-rich and Fe-rich liquids enriched by Zr have relatively good glass-forming ability. When the temperature approaches Tg, the two liquids subsequently undergo LGT. Finally, the microstructure containing two amorphous phases and NCs is obtained, as shown in Fig. 5. The alloy system with the composition at the CuZr-rich side in Fig. 8(a) has a good glass-forming ability [50]. The glass-transition-temperature (Tg) line has an intersection with the binodal line. As a result, there exists a competition between the LLPS and LGT. A monolithic glassy structure may be obtained if the LGT prior to the LLPS.

Concerning the formation of nanoscale structure in (Cu1-xFex)40Zr60 system, it can be attributed to the decrease of the metastable miscibility gap of the (Cu1-xFex)40Zr60 alloy in contrast to that of the (Cu1-xFex)90Zr10 system. As the schematic diagram (Fig. 8(b)), the nucleated droplets, owe to the narrow temperature region Hs-g (=Ts-Tg), have no enough time to grow by diffusion before LGT. On the other hand, the high viscosity (106‒107 poise [51]) and small diffusion coefficient of solute lead to a very small growth rate of the nucleated droplets when the melt temperature is deeply undercooled to approach the glass transition temperature.

3.3. Resistivity behaviour

Due to electical resitivity is sensitive to structural change, measuring resistance change is an effective method to study crystallization. Bulk metallic glasses (BMGs) generally exhibit negative temperature coefficient of resistance (TCR) because the appearance of grain changes the disorder arrangement of atoms as well as the band structure [[52], [53], [54]]. Variation of normalized resistivity (ρT/ρ373K) with temperature of the as-quenched (Cu0.5Fe0.5)40Zr60 alloy at a heating rate of 10 K min-1 are shown in Fig. 9(a). Unlike other Zr-based metallic glasses [52,53], which show continuous decrease of resistivity during crystallization, the resistivity of the (Cu0.5Fe0.5)40Zr60 alloys shows an abnormal increase at the initial stage before the normal declining stage. In addition, the value of resistance relative change is pretty big and this phenomenon is detected for the first time.

Fig. 9.   (a) Variation of nomalized resistivity with temperature and DSC heating curve of (Cu0.5Fe0.5)40Zr60 alloys, (b) Variation of nomalized resistivity with temperature of Zr60Cu40 and Zr75Fe25 alloys.

The DSC curve of as-quenched (Cu0.5Fe0.5)40Zr60 alloys exhibits two exothemic peaks at a heating rate of 20 K min-1 (Fig. 9(a)), which implies a two-step crystallization process. Combine the resistivity and DSC curves, the onset temperature of abnormal increase (693 K) and normal declining (759 K) in resistivity curve are almost consistent with the temparature of crystalliztion (Tx1 = 695 K and Tx2 = 770 K) in DSC curve, which indicates that the change of resistivity is related with first crystallization.

For its structural particularity of the (Cu0.5Fe0.5)40Zr60 alloys, the variation of normalized resistivity (ρT/ρ373K) with temperature of as-quenched Zr60Cu40 and Zr75Fe25 alloys were investigated (Fig. 9(b)). Two distinct resistivity drops occur on the resistance curve of the Zr60Cu40 alloys, which shows a two-step crystallization process. However, the resistance curve of Zr75Fe25 alloys also exihibits an abnormal increase during crystallization. It can be deduced that the abnormal increase in as-quenched (Cu0.5Fe0.5)40Zr60 alloys is related with the crystallization of FeZr-rich matrix.

The crystallization behavior can be evaluated by comparing the activation energies. The most frequently used method for the kinetics of the glass transition and crystallization is based on the Kissinger equation [55,56]:

$ln\frac{T^{2}}{\beta}=\frac{E}{RT}+C$ (4)

where β represents the heating rate (K min-1) and C refers to the constant. R is the gas constant and T stands for Tx or Tp (K). Plots of ln(T2/β) versus 1000/T yield an approximate straight line, whose slope provides the activation energy E. Onset crystallization activation energy Ex1 and crystallization peak activation energy Ep represent the energy of nucleation and grain growth [57,58], respectively. Fig. 10 shows that plots of as-quenched Zr60Cu40, Zr75Fe25 and (Cu0.5Fe0.5)40Zr60 alloys. From the calculated results, it shows that the Ex value of Zr75Fe25 alloys is much smaller, meaning that the crystallization takes place in the FeZr-rich matrix prior to the CuZr-rich particles duing heating the as-quenched (Cu0.5Fe0.5)40Zr60 alloys. What’s more, for the as-quenched Zr75Fe25 alloys, the activation energy Ex is smaller than crystallization peak activation energy Ep, showing that the crystallization of as-quenched Zr75Fe25 alloys is characterized by high population of nucleation and low growth rate. In contrast, for the as-quenched Zr60Cu40 alloys, the Ex value is larger than crystallization peak activation energy Ep.

Fig. 10.   Relationship between ln(T2/β) and 1/T. The subscripts 1, 2 and 3 present Zr75Fe25, Zr60Cu40, and (Cu0.5Fe0.5)40Zr60 alloys, respectively.

To prove the above analysis, the phase evolution during the crystallization was obtained from the XRD patterns after the (Cu0.5Fe0.5)40Zr60 samples were isothermally annealed at 673 K, 713 K and 813 K for 0.5 h as shown in Fig. 11. The diffraction peaks broaden when annealing at 673 K (before abnormal increase of resistivity) for 0.5 h indicates the formation of NCs. There are obvious lattice fringes in the HRTEM image (Fig. 12(a)), which shows that NCs precipitate from FeZr-rich matrix. The formation of NCs result in the formation of nanometer grain boundaries in amorphous matrix and complete long-range ordered structure has been not formed. Finally, the strong electron scattering at the amorphous/ NCs interfaces will result in the increase of resistivity [55].

Fig. 11.   XRD patterns of as-quenched (Cu0.5Fe0.5)40Zr60 alloys annealed at 673 K, 713 K, and 813 K for 0.5 h, respectively.

Fig. 12.   (a) and (b) HRTEM images of (Cu0.5Fe0.5)40Zr60 alloys annealed at 673 K for 0.5 h.

At 713 K (resistivity rising stage) annealed for 0.5 h, the diffraction peaks in XRD patterns obviously broaden and no new phases form, which shows a large of NCs start to nucleate and grow slowly compared with XRD patterns of annealing at 673 K for 0.5 h. This is consistent with calculated activation energy. Nucleation of a large of NCs with slow growth will bring lots of nanometer grain boundaries. Therefore, the resistivity continues to increase and the maximum of the resistance relative change was about 12.5%, which is bigger than that in literature [59] and Zr75Fe25 alloys. It could be that there is an interface between the amorphous nanoparticles and amorphous matrix (as shown in Fig. 12(b)) in addition to amorphous/NCs interfaces. It may also enhance the electron scattering, which results in a large increase of resistivity. At 813 K (after normal decline), peaks belonging to NCs vanish and new peaks appear (ie., Zr2(Cu,Fe) phase). It indicates that NCs finally grow up and residual amorphous phase transforms to a new stable phase completely. Because NCs grew up and the stable crystal structure formed, nanometer grain boundaries vanished and resistivity decreased finally.

4. Conclusion

The microstructural evolution and competitive mechanism of phase formation in the rapidly solidified Cu-Fe-Zr system were studied experimentally and by thermodynamic calculations. The liquid-liquid phase separation into the Cu-rich and Fe-rich liquids may occur during the Cu-Fe-Zr alloy melt cooling in the metastable miscibility gap. The microstructure from interconnected type to particle-isolated type is tunable by modifying the atomic ratio of Cu to Fe in the (Cu1-xFex)90Zr10 alloy. Combining the thermodynamic calculations with DTA analyses indicates that the dome height of the miscibility gap decreases with the increase of Zr content. The structure of the rapidly solidified (Cu1-xFex)40Zr60 alloy is determined by the competition among the liquid-liquid, liquid-solid, and liquid-glass transitions. For the (Cu0.5Fe0.5)40Zr60 alloy, the nanoscale liquid-liquid phase separation and then liquid-glass transition can take place during the rapid solidification and result in a heterogeneous structure with glassy Fe-rich matrix embedded by glassy Cu-rich nanoparticles. The characteristic size of the Cu-rich nanoparticle is 2-8 nm, and the population density and volume fraction of the particles are estimated to be 5.37 × 1024 m-3 and 10.4% ± 0.2%. Moreover, the electrical property of the as-quenched (Cu0.5Fe0.5)40Zr60 alloy samples was investigated. It is of interest in an abnormal increase of electrical resistivity with the increase of temperature at the stage of crystallization. This may be attributed that the formation of high population density nanocrystals with slow growth in the glassy Fe-rich matrix results in the enhancement of electron scattering.

Acknowledgment

This work was supported by the National Natural Science Foundation of China (Nos. 51774264, 51574216, 51974288 and 51374194) and the Natural Science Foundation of Liaoning Province of China (No. 2019-MS-332).


Reference

[1] L. Ratke, S. Diefenbach, Mater. Sci. Eng. R 15 (1995) 263-347.

[Cited within: 1]     

[2] J. He, C. Xing, J. Zhao, L. Zhao, J. Mater. Sci. Technol. 26(2010) 136-140.

[Cited within: 1]     

[3] N. Yan, Z.Y. Hong, D.L. Geng, W.L. Wang, B. Wei, J. Alloys. Compd. 544(2012) 6-12.

[Cited within: 1]     

[4] L.Y. Chen, J.Q. Xu, H. Choi, H. Konishi, S. Jin, X.C. Li, Nat. Commun. 5(2014) 3879.

DOI      URL      PMID      [Cited within: 1]      Abstract

Effective control of phase growth under harsh conditions (such as high temperature, highly conductive liquids or high growth rate), where surfactants are unstable or ineffective, is still a long-standing challenge. Here we show a general approach for rapid control of diffusional growth through nanoparticle self-assembly on the fast-growing phase during cooling. After phase nucleation, the nanoparticles spontaneously assemble, within a few milliseconds, as a thin coating on the growing phase to block/limit diffusion, resulting in a uniformly dispersed phase orders of magnitude smaller than samples without nanoparticles. The effectiveness of this approach is demonstrated in both inorganic (immiscible alloy and eutectic alloy) and organic materials. Our approach overcomes the microstructure refinement limit set by the fast phase growth during cooling and breaks the inherent limitations of surfactants for growth control. Considering the growing availability of numerous types and sizes of nanoparticles, the nanoparticle-enabled growth control will find broad applications.
[5] Q. Sun, H. Jiang, J. Zhao, J. He, Acta Mater. 129(2017) 321-330.

[Cited within: 1]     

[6] W. Li, H.X. Jiang, L.L. Zhang, S.X. Li, J. He, J.Z. Zhao, F. Ai, Scr. Mater. 162(2019) 426-431.

[Cited within: 1]     

[7] S.I. Hong, J.S. Song, Metall. Mater. Sci. 32(2001) 985-991.

[Cited within: 1]     

[8] N.D. Stepanov, A.V. Kuznetsov, G.A. Salishchev, N.E. Khlebova, V.I. Pantsyrny, Mater. Sci. Eng. A 564 (2013) 264-272.

[Cited within: 1]     

[9] C. Biselli, D.G. Morris, Acta Mater. 44(1996) 493-504.

[Cited within: 1]     

[10] K.M. Liu, D.P. Lu, H.T. Zhou, Z.B. Chen, A. Atrens, L. Lu, Mater. Sci. Eng. A 584 (2013) 114-120.

[Cited within: 1]     

[11] A. Bachmaier, M. Kerber, D. Setman, R. Pippan, Acta Mater. 60(2012) 860-871.

DOI      URL      PMID      [Cited within: 1]      Abstract

Fully dense bulk nanocomposites have been obtained by a novel two-step severe plastic deformation process in the immiscible Fe-Cu system. Elemental micrometer-sized Cu and Fe powders were first mixed in different compositions and subsequently high-pressure-torsion-consolidated and deformed in a two-step deformation process. Scanning electron microscopy, X-ray diffraction and atom probe investigations were performed to study the evolving far-from-equilibrium nanostructures which were observed at all compositions. For lower and higher Cu contents complete solid solutions of Cu in Fe and Fe in Cu, respectively, are obtained. In the near 50% regime a solid solution face-centred cubic and solid solution body-centred cubic nanograined composite has been formed. After an annealing treatment, these solid solutions decompose and form two-phase nanostructured Fe-Cu composites with a high hardness and an enhanced thermal stability. The grain size of the composites retained nanocrystalline up to high annealing temperatures.
[12] Y.Y. Chuang, R. Schmid, Y.A. Chang, Metall. Mater. Trans. A 15 (1984) 1921-1930.

[Cited within: 1]     

[13] Q. Chen, Z.P. Jin, Metall. Mater. Trans. A 26 (1995) 417-426.

[Cited within: 1]     

[14] J. He, J.Z. Zhao, L. Ratke, Acta Mater. 54(2006) 1749-1757.

[Cited within: 2]     

[15] C.P. Wang, X.J. Liu, I. Ohnuma, R. Kainuma, K. Ishida, Science 297 (2002) 990-993.

DOI      URL      PMID      [Cited within: 2]      Abstract

The egg-type core microstructure where one alloy encases another has previously been obtained during experiments in space. Working with copper-iron base alloys prepared by conventional gas atomization, we were able to obtain this microstructure under gravity conditions. The minor liquid phase always formed the core of the egg, and it sometimes also formed a shell layer. The origin of the formation of this core microstructure can be explained by Marangoni motion on the basis of the temperature dependence of the interfacial energy, which shows that this type of powder can be formed even if the cooling rate is very high.
[16] R.P. Shi, C.P. Wang, D. Wheeler, X.J. Liu, Y. Wang, Acta Mater. 61(2013) 1229-1243.

[Cited within: 2]     

[17] Y. Nakagawa, Acta Metall. 6(1958) 704-711.

[Cited within: 1]     

[18] S.P. Elder, A. Munitz, G.J. Abbaschian, Mater. Sci. Forum 50 (1989) 137-150.

[Cited within: 1]     

[19] F.P. Dai, C.D. Cao, B.B. Wei, Sci. China Ser. G-Phys.Mech. Astron. 50(2007) 509-518.

[Cited within: 1]     

[20] A. Munitz, A. Venkert, P. Landau, M.J. Kaufman, R. Abbaschian, J. Mater. Sci. 47(2012) 7955-7970.

[Cited within: 1]     

[21] J.T. Zhang, X.C. Cui, Y.H. Wang, Int. J. Cast. Metals Res. 31(2018) 87-92.

[Cited within: 1]     

[22] S. Liu, J. Jie, B. Dong, Z. Guo, T. Wang, T. Li, Mater. Des. 156(2018) 71-81.

[Cited within: 2]     

[23] S. Liu, J. Jie, Z. Guo, G. Yin, T. Wang, T. Li, J. Alloys. Compd. 742(2018) 99-106.

[Cited within: 2]     

[24] J. He, J.Z. Zhao, J. Mater. Sci. Technol. 21(2005) 759-762.

[Cited within: 1]     

[25] Z.C. Xia, W.L. Wang, S.B. Luo, B. Wei,J. Appl. Phys. 117(2015), 054901.

[Cited within: 1]     

[26] C.P. Wang, X.J. Liu, R.P. Shi, C. Shen, Y. Wang, I. Ohnuma, R. Kainuma, K. Ishida,Appl. Phys. Lett. 91(2007), 141904.

DOI      URL     

[27] C.P. Wang, X.J. Liu, Y. Takaku, I. Ohnuma, R. Kainuma, K. Ishida, Metall. Mater. Trans. A 35 (2004) 1243-1253.

DOI      URL     

[28] M. Nakamoto, Y. Tanaka, S. Tagaki, T. Tanaka, T. Yamamoto, ISIJ Int. 53(2013) 551-556.

DOI      URL      [Cited within: 1]     

[29] J.S. Song, S.I. Hong, Y.G. Park, J. Alloys. Compd. 388(2005) 69-74.

[Cited within: 1]     

[30] S.B. Luo, W.L. Wang, Z.C. Xia, Y.H. Wu, B. Wei, Appl. Phys. A 119 (2015) 1003-1011.

[Cited within: 1]     

[31] S.B. Luo, W.L. Wang, Z.C. Xia, B.B. Wei, Trans. Nonferrous Met. Soc. China 26 (2016) 2762-2769.

[Cited within: 1]     

[32] B. Chen, J. He, Y.Y. Xi, X.F. Zeng, I. Kaban, J.Z. Zhao, H.R. Hao, J. Hazard. Mater.364(2019) 388-395.

DOI      URL      PMID      [Cited within: 1]      Abstract

Fe-rich + LCu,Pb-rich, LCu,Pb-rich→LCu-rich + LPb-rich and LPb-rich→SCu-dendritical + L'Pb-rich produces four immiscible Fe-rich, Cu-rich, Cu-dendritical and Pb-rich substances. The separation rate between these substances can reach more than 96% in a super-gravity field of G = 1000g. Other metals selectively distribute in the four substances. The Fe-rich substance collects Cr, Co, Ni and Si. Almost all of Au and Ag are trapped in the Cu-rich and Cu-dendritical substances. The low-melting-point metals, i.e. Bi, Cd, In and Sn, are located in the Pb-rich substance. This work provides a green shortcut for efficiently separating and recycling overall metals in WPCBs.]]>
[33] B. Chen, J. He, X.J. Sun, J.Z. Zhao, H.X. Jiang, L.L. Zhang, H.R. Hao, Acta Metall. Sin. 55(2019) 751-761 (in Chinese).

[Cited within: 1]     

[34] C.P. Wang, X.J. Liu, I. Ohnuma, R. Kainuma, K. Ishida, J. Phase Equilib. 23(2002) 236-245.

[Cited within: 1]     

[35] W.L. Wang, L. Hu, S.B. Luo, L.J. Meng, D.L. Geng, B. Wei, Intermetallics 77 (2016) 41-45.

[Cited within: 1]     

[36] R. Wang, W. Chen, J. Zhong, L. Zhang, J. Mater. Sci. Technol. 34(2018) 1791-1798.

DOI      URL      [Cited within: 1]     

[37] J. He, N. Mattern, I. Kaban, F.P. Dai, K.K. Song, Z.J. Yan, J.Z. Zhao, D.H. Kim, J. Eckert, J. Alloys. Compd. 618(2015) 795-802.

[Cited within: 1]     

[38] J. He, I. Kaban, N. Mattern, K.K. Song, B.A. Sun, J.Z. Zhao, D.H. Kim, J. Eckert, A.L. Greer, Sci. Rep. 6(2016) 25832.

DOI      URL      PMID      [Cited within: 1]      Abstract

At room temperature, plastic flow of metallic glasses (MGs) is sharply localized in shear bands, which are a key feature of the plastic deformation in MGs. Despite their clear importance and decades of study, the conditions for formation of shear bands, their structural evolution and multiplication mechanism are still under debate. In this work, we investigate the local conditions at shear bands in new phase-separated bulk MGs containing glassy nanospheres and exhibiting exceptional plasticity under compression. It is found that the glassy nanospheres within the shear band dissolve through mechanical mixing driven by the sharp strain localization there, while those nearby in the matrix coarsen by Ostwald ripening due to the increased atomic mobility. The experimental evidence demonstrates that there exists an affected zone around the shear band. This zone may arise from low-strain plastic deformation in the matrix between the bands. These results suggest that measured property changes originate not only from the shear bands themselves, but also from the affected zones in the adjacent matrix. This work sheds light on direct visualization of deformation-related effects, in particular increased atomic mobility, in the region around shear bands.
[39] S. Michalik, K. Saksl, P. Sovák, K. Csach, J.Z. Jiang, J. Alloys. Compd. 478(2009) 441-446.

DOI      URL      [Cited within: 1]     

[40] A. Takeuchi, A. Inoue, Mater. Trans. 46(2005) 2817-2829.

DOI      URL      [Cited within: 1]     

[41] I. Kaban, P. Jovari, M. Stoica, N. Mattern, J. Eckert, W. Hoyer, B. Beuneu,J. Phys. Condes. Matter 22 (2010), 404208.

DOI      URL      PMID      [Cited within: 1]      Abstract

The structure of Zr(60)Cu(20)Fe(20) metallic glass has been studied with high-energy x-ray diffraction, neutron diffraction and extended x-ray absorption spectroscopy and modelled with the reverse Monte Carlo simulation technique. It is found that Cu and Fe atoms prefer Zr as a nearest neighbour. The mean interatomic distance between Cu/Fe and Zr atoms in the glass is remarkably shorter than the sum of the respective atomic radii. The coordination numbers for Cu/Fe-Cu/Fe pairs are very close to each other, suggesting a regular distribution of Cu and Fe atoms in the Zr(60)Cu(20)Fe(20) metallic glass.
[42] S. Sengul, M. Celtek, U. Domekeli, Vacuum 136 (2017) 20-27.

DOI      URL      [Cited within: 1]     

[43] X.H. Wang, A. Inoue, F.L. Kong, S.L. Zhu, M. Stoica, I. Kaban, C.T. Chang, E. Shalaan, F. Al-Marzouki, J. Eckert, Acta Mater. 116(2016) 370-381.

DOI      URL      [Cited within: 1]     

[44] N. Zou, H.J. Lu, P.W. Gu, J.Y. Wang, X.K. Feng, J.Y. Shen, Y.L. He, L. Li, X.G. Lu, Calphad 64 (2019) 175-184.

DOI      URL      [Cited within: 1]     

[45] L.A. Dreval, P.G. Agraval, M.A. Turchanin, Metall. Mater. Trans. B 46 (2015) 2234-2245.

DOI      URL      [Cited within: 1]     

[46] C.P. Guo, Z.M. Du, C.R. Li, B.L. Zhang, M. Tao, Calphad 32 (2008) 637-649.

DOI      URL      [Cited within: 1]      Abstract

AbstractThe Fe–Zr and Al–Fe–Zr systems were critically assessed by means of the CALPHAD technique. The solution phases, liquid, face-centered cubic, body-centered cubic and hexagonal close-packed, were described by the substitutional solution model. The compounds with homogeneity ranges, hex.- Fe2Zr, Fe2Zr, FeZr2 and FeZr3 in the Fe–Zr system, were described by the two-sublattice model in formulas such as hex.- Fe2(Fe,Zr), (Fe,Zr)2(Fe,Zr), (Fe,Zr)Zr2 and (Fe,Zr)(Fe,Zr)3 respectively. The compounds AlmZrn except Al2Zr in the Al–Zr system were treated as line compounds (Al,Fe)mZrn in the Al–Fe–Zr system. The compounds FeZr2 and FeZr3 in the Fe–Zr system were treated as (Al,Fe,Zr)Zr2 and (Al,Fe,Zr)(Fe,Zr)3 in the Al–Fe–Zr system, respectively. All compounds in the Al–Fe system and hex.- Fe2Zr in the Fe–Zr system have no solubilities of the third components Zr or Al, respectively, in the Al–Fe–Zr system. The ternary compounds λ1 with C14 structure and λ2 with C15 structure in the Al–Fe–Zr system were treated as λ1- (Al,Fe,Zr)2(Fe,Zr) with Al2Zr in the Al–Zr system and λ2- (Al,Fe,Zr)2(Fe,Zr) with Fe2Zr in the Fe–Zr system, respectively. And the ternary compounds τ1, τ2 and τ3 in the Al–Fe–Zr system were treated as (Al,Fe)12Zr, Fe(Al,Zr)2Zr6 and Fe7Al67Zr26, respectively. A set of self-consistent thermodynamic parameters of the Al–Fe–Zr system was obtained.]]>
[47] H.M. Hsiao, S.M. Liang, R. Schmid-Fetzer, Y.W. Yen, Calphad 55 (2016) 77-87.

DOI      URL      [Cited within: 1]     

[48] J.T. Zhang, X.C. Cu, Y.H. Wang, Int. J. Cast Metal.Res. 31(2018) 87-92.

[Cited within: 1]     

[49] S. Curiotto, L. Battezzati, E. Johnson, N. Pryds, Acta Mater. 55(2007) 6642-6650.

DOI      URL      [Cited within: 1]      Abstract

AbstractThis work deals with the effect of Ni addition on the liquid–liquid phase separation in the Cu–Co system which displays a liquid metastable miscibility gap. Cu–Co–Ni samples with different compositions have been processed by differential scanning calorimetry using a fluxing technique to precisely determine transformation temperatures. The ternary phase diagram has been evaluated by Calphad, extrapolating the binary systems and optimizing ternary parameters using the experimental results of the present work. Ni additions reduce the demixing temperature, bringing the binodal line below the peritectic. The mechanisms of demixing and formation of microstructures have been clarified for low and high Ni content.]]>
[50] T.L. Wang, B.X. Liu, J. Alloys. Compd. 481(2009) 156-160.

DOI      URL      [Cited within: 1]     

[51] E. Bakke, R. Busch, W.L. Johnson, Appl. Phys. Lett. 67(1995)3260-3262.

DOI      URL      [Cited within: 1]     

[52] S.J. Chung, K.T. Hong, M.R. Ok, J.K. Yoon, G.H. Kim, Y.S. Ji, B.S. Seong, K.S. Lee, Scr. Mater. 53(2005) 223-228.

DOI      URL      [Cited within: 2]     

[53] O. Haruyama, T. Miyazawa, J. Saida, A. Inoue, Appl. Phys. Lett. 79(2001) 758-760.

DOI      URL      [Cited within: 2]     

[54] Y.S. Ji, S.J. Chung, M.R. Ok, K.T. Hong, J.Y. Suh, J.W. Byeon, J.K. Yoon, K.H. Lee, K.S. Lee, Mater. Sci. Eng. A 449 (2007) 521-525.

[Cited within: 1]     

[55] H.E. Kissinger, J. Res. Natl. Bur. Stand. 57(1956) 217-221.

DOI      URL      [Cited within: 2]     

[56] H.E. Kissinger, Anal. Chem. 29(1957) 1702-1706.

DOI      URL      [Cited within: 1]     

[57] F.X. Qin, H.F. Zhang, B.Z. Ding, Z.Q. Hu, Intermetallics 12 (2004) 1197-1203.

[Cited within: 1]     

[58] H.D. Wei, Q.H. Bao, C.X. Wang, W.S. Zhang, Z.Z. Yuan, X.D. Chen, J. Non-Cryst. Solids 354 (2008) 1876-1882.

[Cited within: 1]     

[59] B.B. Liu, N.N. Zuo, F. Ye, Mater. Lett. 171(2016) 285-288.

[Cited within: 1]     

/