Next Article in Journal
Ignimbrites Related to Neogene Volcanism in the Southeast of the Iberian Peninsula: An Experimental Study to Establish Their Pozzolanic Character
Previous Article in Journal
Effect of Surface Dispersion of Fe Nanoparticles on the Room-Temperature Flash Sintering Behavior of 3YSZ
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Annealing on the Interface and Properties of Pd/Al Composite Wires

1
State Key Laboratory of Nonferrous Metals and Processes, China GRINM Group Co., Ltd., Beijing 100088, China
2
GRIMAT Engineering Institute Co., Ltd., Beijing 101407, China
3
General Research Institute for Nonferrous Metals, Beijing 100088, China
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(4), 1545; https://doi.org/10.3390/ma16041545
Submission received: 5 January 2023 / Revised: 25 January 2023 / Accepted: 6 February 2023 / Published: 13 February 2023
(This article belongs to the Topic Metal Matrix Composites: Recent Advancements)

Abstract

:
This paper investigates the changes in the interface organization and properties of 0.10 mm Pd/Al composite wires annealed at different temperatures. The optimum comprehensive performance of the material was obtained after annealing at 300 °C for 120 s. Its tensile strength, conductivity and elongation are 140.61 MPa, 46.82%IACS and 14.89%, respectively. Scanning electron microscopy (SEM) and transmission electron microscopy (TEM) were used to observe the intermetallic compounds on the interface. The annealing temperature and the formation heat of intermetallic compounds determine the categories and evolution of intermetallic compounds. When the thickness of the intermetallic layer is more than 1 μm, it has a serious effect on the electrical conductivity and elongation of the materials.

1. Introduction

Bi-metallic composite wire [1] is a composite wire with a core and an outer layer made up of two different metals or alloys. Bimetallic composite wire has excellent properties of two materials, can achieve the performance that a single material cannot, and is widely used in all walks of life. For example: copper-clad aluminium [2,3,4,5] composite wire has both the good electrical conductivity of copper and the light weight and good ductility of aluminium, and is widely used in high frequency transmission cables and other fields. Silver-clad aluminium [6,7] has excellent electrical conductivity, good contact properties and low density, and is used as aerospace wire and waveguides. Pd/Al composite wire [8] is an aluminium-based composite bimetallic wire with the advantages of low density, light weight and the good electrical conductivity of aluminium. As a reactive thermal fuse, it can be supplied with energy from AC, DC or capacitive discharge to heat the composite wire. As the rate of Pd/Al diffusion increases, a large number of intermetallic compounds are generated at the interface, and an immediate exothermic reaction occurs, generating high temperatures that can directly detonate the explosive or ignite the solid fuel. Because of its short reaction time, the availability of a constant current to control the reaction process and the high heat release during the reaction (about 1368.19 J/g), it can directly meet the requirements of short ignition response time, small ignition peaks and easy control of the ignition time for solid rocket motors [9,10,11,12]. It has the characteristics of having less substance discharged during the reaction, heat transferred mainly by conduction, the reaction does not need oxygen, and it can react under vacuum and inert gas. It is suitable for detonation underwater, underground, at high altitude and in a vacuum. It is used as a detonating wire in fields such as torpedoes and new electronic detonators [13,14].
The ignition reaction of Pd/Al composite wires mainly depends on the reaction heat during the formation of intermetallic compounds. However, the production of intermetallic compounds not only affects the basic mechanical and electrical properties of the composite wire, but also plays a key role in the effectiveness of the application of the composite wire. According to the Al-Pd phase diagram [15], it can be seen that the Pd/Al phase diagram is complex, with the existence of nine different equilibrium phases depending on the elemental content, namely Al4Pd, Al3Pd, Al21Pd8, Al3Pd2, AlPd, Al3Pd5, Al2Pd5 and AlPd2 phases. Many researchers have carried out correlation analyses for the performance of Pd/Al composite wires at different temperatures [16,17] and ignition responses [18]. Howard [19] found that PdAl3 was the initial phase of Pd/Al composites at 250 °C and 300 °C. The presence of the Al3Pd2, AlPd, Al3Pd5 and AlPd2 phases was observed in Pd/Al composites heat treated at 200 °C to 500 °C by KÖster and Hon [20]. The presence of PdAl3 and PdAl4 phases was also observed. It is believed that β-AlPd is the first to be formed in all samples and is accompanied by the formation of the Al3Pd2 phase. Nastasi and Hung [21] found PdAl4 as the initial phase by in situ annealing, and with further heat treatment, the PdAl4 and PdAl3 phases were obtained. However, these studies rely on scanning electron microscopy, energy spectral analysis, X-ray diffraction, etc., and cannot make a precise determination of the crystal structure of the phases.
According to the Cu-Al diffusion couple [22,23], after holding at a certain temperature for a period of time, a series of different intermetallic compounds will appear on the interface. The heat treatment system is the main factor affecting the bonding of Cu-Al. However, the growth of intermetallic compounds in Pd/Al composites after annealing at different temperatures remains unclear. Therefore, in order to ensure that the Pd/Al composites obtained meet the application requirements, in the preparation and processing of composite materials, it is necessary to have an in-depth understanding of the types of compounds on the interface and the sequence of compounds generated under the specific preparation process, so as to better control them.
In this study, a combination of transmission electron microscopy and scanning electron microscopy was used to observe the intermetallic compound layer changes of Pd/Al composite wires after annealing. The aim is to investigate the generation of phases during annealing on the Pd/Al interface, and to investigate the effect of the intermetallic compounds on the overall properties of Pd/Al composite wires. This work seeks a deeper understanding of the phase formation and transformation, the relationship between different intermetallic compounds and the properties of Pd/Al composite wires.

2. Experimental Methods

Palladium tubes with an outer diameter of 32.00 mm and an inner diameter of 25.80 mm and solid aluminium rods with an outer diameter of 25.80 mm were prepared using 99.999% pure palladium and pure aluminium in a volume ratio of 35:65. After drawing, the Pd/Al composite wire diameter (D) was 0.10 mm.
Next came annealing of the 0.10 mm Pd/Al composite wire at 200 °C, 250 °C, 300 °C and 350 °C for 5 s, 10 s, 30 s, 60 s, 120 s, 300 s, 900 s, 1800 s, 3600 s, 7200 s and 10,800 s, respectively. The TH-8203 S tensile testing machine was used to carry out the tensile test, in which the length of the gauge was 200 mm and the tensile speed was 50 mm/min. The conductivity was measured by TX-300 conductivity tester (Xiamen, China). In order to ensure the accuracy and reliability of the data, five sets of data were tested for each sample in each regime, and the results were averaged. The data of the sample in its initial state without heat treatment were also tested for comparison.
The annealed samples were inlaid using a cold inlay on a cross-section of palladium-clad aluminium wire, after coarse grinding, fine grinding and polishing. The intermetallic compounds at the material interface were observed by scanning electron microscopy (SEM) and transmission electron microscopy (TEM), where TEM samples were obtained by focused ion beam (FIB) preparation. The component content was analysed by an energy dispersive spectrometer (EDS). The thickness of the intermetallic layer was measured by energy spectrum results.

3. Results and Discussion

3.1. Interface Analysis

Figure 1 shows an SEM image of the cross-section of an unheated treated Pd/Al composite wire. At low magnification (as shown in Figure 1a), it can be seen that the material has a clean cross-section with no impurities or oxides present (The outer white part is Pd and the inner black part is Al). Due to the different hardness of Pd and Al, the Pd/Al interface shows a cogwheel shape after intense cold deformation. Under high magnification (as shown in Figure 1b), it can be seen that the interface is tightly bonded and no holes exist, indicating that the interface of the Pd/Al composite wire prepared by the cold-drawing method is well-bonded.
Figure 2 shows SEM images of Pd/Al composite wires annealed at 200 °C for different times, with Al in the darker areas and Pd in the lighter areas due to the difference in lining degree. As can be seen from the figure, when the annealing time is relatively short (as shown in Figure 2a), no intermetallic compound layer is observed at the interface of the material. When the annealing time reached 60 s (as shown in Figure 2b), a thin layer of the intermetallic compound appeared at the interface of the two phases of the material. With a further increase in annealing time, the intermetallic compound layer of the material further thickened, increasing to an average of about 0.85 μm at 7200 s, and when the annealing reached 10,800 s (as shown in Figure 2d), the thickness of the interfacial layer reached about 1.1 μm.
Figure 3 shows TEM images of Pd/Al composite wires annealed at 200 °C for 5 s and 10,800 s. Figure 3a,e show high-angle annular dark-field (HAADF) diagrams for 5 s and 10,800 s annealed at 200 °C. It can be seen that at an annealing time of 5 s, a very thin intermetallic compound layer appears at the interface. Through high-resolution images according to fast Fourier transform (FFT), this was identified as a hexagonal Al3Pd2 phase in the (0 2 −1) direction (as shown in Figure 3c). Position 4 in Figure 3f shows the polycrystalline ring, according to the selected area electron diffraction (SAED) result, confirmed as a hexagonal Al3Pd2 structure. Figure 3g shows a selected diffraction electron image at position 5. Position 5 reveals a hexagonal Al3Pd2 phase in the (1 −1 2) direction (black quadrilateral in Figure 3g) and a tetragonal Al21Pd8 (red quadrilateral in Figure 3g) phase in the (1 −1 0) direction. This indicates that at 200 °C, the material interface is dominated by the growth of the Al3Pd2 phase and the formation of a small portion of the Al21Pd8 phase.
Figure 4 shows SEM images of Pd/Al composite wires annealed at 250 °C for different times. A thin interfacial layer can be seen at 250 °C when the annealing time reaches 30 s (as shown in Figure 4b) at the interface of the material. With increasing annealing time, the intermetallic compound layer at the interface of the material thickens rapidly with an average thickness of about 0.67 μm when the annealing time is increased to 1800 s (as shown in Figure 4c). It is worth noting that with a further increase in annealing time, the thickness of the interfacial layer of the material does not vary significantly, with annealing time of 3600 s (as shown in Figure 4d) and 10,800 s (as shown in Figure 4e), respectively, both being around 1.163 μm. It indicates that, in the early stages of annealing, the mutual diffusion of Pd and Al is rapid due to direct contact, but as the intermetallic compound layer is created, the diffusion drive decreases, resulting in the slow growth of the interfacial layer. This phenomenon was also seen in the samples annealed at 200 °C.
Figure 5 shows TEM images of annealing at 250 °C for 5 s and 10,800 s. Figure 5a shows that, at this temperature, the material generates intermetallic compounds on the interface at the early stage of annealing that are extremely small and discontinuous, with an intermetallic compound thickness of about 30 nm, and the generated intermetallic compound is a hexagonal Al3Pd2 phase in the (0 0 1) direction (as shown in Figure 5c). At an annealing time of 10,800 s, two different intermetallic compound layers appeared at the interface of the material, in the region of position 2 and position 3 in Figure 5e, with the hexagonal Al3Pd2 phase (black quadrilateral in Figure 5f,g) and the tetragonal Al21Pd8 phase (red quadrilateral in). According to the diffusion phase ratio, only one phase can be present in the diffusion region for both metals. The EDS results suggest that the region at position 2 is the Al3Pd2 phase region and position 3 is the Al21Pd8 phase region. The presence of the Al21Pd8 phase was not directly observed at the beginning of the annealing process, indicating that the growth of the Al21Pd8 phase was backward. Due to the thin thickness of the Al21Pd8 phase, it was not observed in Figure 4.
Figure 6 shows SEM images of the Pd/Al composite wire annealed at 300 °C for different times. When the annealing time reaches 30 s (as shown in Figure 6b), a grey intermetallic compound layer appears at the Pd/Al interface bond (red arrow area in the figure). With a further increase in annealing time, by 900 s (as shown in Figure 6c), the presence of the intermetallic compound layer at the interface of the material is clearly visible at the interface, where the thickness of this compound layer is approximately 0.52 μm. When the annealing time reached 10,800 s (as shown in Figure 6f), the interfacial layer of the material was further thickened to approximately 3.15 μm. The higher diffusion drive allows further growth of the interfacial layer as the temperature increases compared to 250 °C.
Figure 7 shows TEM images of annealing at 300 °C for 5 s and 10,800 s. It can be seen that, at an annealing time of 5 s, a discontinuous intermetallic compound layer appears at the interface, which, after EDS calibration and FFT transformation at high resolution, is identified as a hexagonal Al3Pd2 phase in the (0 1 0) direction, with an intermetallic compound layer thickness of approximately 20 nm. At an annealing time of 10,800 s, it was observed that the intermetallic compound layer of the material was only one layer of the material, consisting of a large number of fine grains, and the material showed a polycrystalline ring structure, which was considered to be a hexagonal Al3Pd2 structure by calibration.
Figure 8 shows SEM images of Pd/Al composite wires annealed at 350 °C for different times. It can be seen that, at an annealing time of 5 s (as shown in Figure 8a), a white continuous intermetallic compound layer is observed at the interface of the material, and as the annealing time increases by 60 s (as shown in Figure 8a), the intermetallic compound layer of the material grows to about 0.48 μm. At an annealing time of 300 s (as shown in Figure 8d), the thickness of the interfacial layer of the material increased to 1.01 μm. This indicates that the interfacial layer grew faster and faster as the annealing temperature increased. When the annealing time came to 3600 s (as shown in Figure 8e), an off-white layer appeared at the interface, and it is believed that a new intermetallic compound was produced at this time. At an annealing time of 10,800 s (as shown in Figure 8f), the new intermetallic compound grows and appears as a white area in the diagram.
Figure 9 shows TEM images of annealing at 350 °C for 5 s and 10,800 s. At an annealing time of 5 s, a continuous and homogeneous layer of the intermetallic compound appears at the interface of the material with an interfacial layer thickness of approximately 7.3 nm. Based on the EDS results at position 1 (Figure 9f) and its Fourier transform image (shown in Figure 9c), it is considered that an orthogonal Al3Pd5 phase in the (1 0 −2) direction is generated at this point. As the annealing time was increased to 10,800 s (shown in Figure 9e), two compound layers appeared at the interface of the material, where the thin layer showed crescent-shaped growth, and the combination of EDS results (shown in Figure 9h) and SAED results showed that the intermetallic compound layers at positions 2 and 3 were the Al3Pd2 and AlPd layers, respectively.
The interfacial intermetallic compounds generated by annealing the material under different conditions are shown in Table 1. An intermetallic compound Al3Pd2 phase was formed at the interface at the beginning of annealing at 200–300 °C. At 350 °C, the Al3Pd5 phase was formed. However, the Al3Pd5 phase is a high temperature phase that is stable above 615 °C. Traces of the Al3Pd5 phase were only observed above 600 °C by A.S. Ramos [16]. A similar situation also appears in Cu/Al composite wires [24]. This is due to the fact that Al is in full contact with Pd and the reaction diffuses at a higher temperature, resulting in a higher instantaneous heat release and, thus, a very high temperature. On the other hand, due to the high solid solution of Al in Pd and the very low solid solution of Pd in Al, it is easier for Al to diffuse into Pd and preferentially reach a concentration suitable for the formation of the Al3Pd5 phase. As the annealing time increases, the diffusion of atoms slows down, the temperature decreases and stabilises and the Al3Pd5 phase transforms into the Al3Pd2 and AlPd phases. After prolonged annealing, the complex effects of external heat transfer, exothermic generation of pre-metallic compounds and the diffusion rates of Al and Pd lead to the generation of different phases at the interfaces annealed at different temperatures. After annealing at 200 °C, 250 °C and 300 °C for 10,800 s, the intermetallic compound at the interface is the Al3Pd2 and Al21Pd8 phases, and after annealing at 350 °C for 10,800 s, the intermetallic compound at the interface is the Al3Pd2 phase.

3.2. Mechanical and Conductive Properties

Figure 10 shows the variation of tensile strength, elongation and electrical conductivity of the Pd/Al composite wires at different temperatures with annealing time. The mechanical and electrical properties of the materials are seriously affected by intermetallic compound layers [19,20]. It can be seen that the initial tensile strength, elongation and electrical conductivity of the materials are 269.91 MPa, 1.22% and 48.08% IACS. When the annealing temperature is 200 °C, the tensile strength and electrical conductivity of the composite wire show a smooth and then decreasing trend, while the elongation shows a smooth and then increasing trend. The inflection point occurs in the time range of 3600 s–7200 s. The reduction in tensile strength and the increase in elongation are mainly due to the recovery of Pd and the recrystallisation of Al. The decrease in electrical conductivity, on the other hand, is due to the growth of interfacial intermetallic compounds.
As the annealing temperature increases, the annealing time used for the inflection point to occur becomes shorter, and the decrease in tensile strength and electrical conductivity becomes more significant. The elongation of the composite wire reaches its peak at 200 °C annealed for 3600 s, 250 °C annealed for 1800 s, 300 °C annealed for 120 s and 350 °C annealed for 120 s, and the tensile strength and electrical conductivity remain at a high level when the thickness of the interfacial layer of the material is all around 1 μm. As the annealing time increases further and the thickness of the interfacial layer continues to increase, the electrical conductivity and elongation of the material decreases sharply.
There is a similar effect of intermetallic compound layers at the copper-clad aluminium interface on material elongation [25,26,27]. This is due to the tendency of the thick intermetallic compound layer to break up during stretching [28,29], and to fail to co-ordinate the simultaneous deformation of Pd/Al, leading to interface separation and severely reducing the elongation of the wires. It can therefore be assumed that 1 μm is the critical point at the interface where the intermetallic compound layer affects the performance of the composite wire. The overall properties of the material were optimal when the material was annealed at 300 °C for 120 s, at which point the tensile strength, elongation and electrical conductivity were 140.61 MPa, 14.89% and 46.82% IACS, respectively.
The conductivity of the material was 48.28% IACS and 45.86% IACS when annealed at 300 °C for 5 s and 350 °C for 5 s, respectively, at which time the interfacial phases were the Al3Pd2 and Al3Pd5 phases, but the thickness of the Al3Pd2 phase (20 nm) was less than that of the Al3Pd5 phase (7.8 nm). Considering the better recovery of Pd when annealed at 350 °C, it suggests that the negative effect of the Al3Pd5 phase on the conductivity is greater than that of the Al3Pd2 phase.

4. Conclusions

It was shown in this study that the intermetallic compound on the interface of the Pd/Al composite wire is in the Al3Pd2 phase at the beginning of annealing at 200–300 °C. At 350 °C, a high temperature stable Al3Pd5 phase is generated. Due to the full contact between Al and Pd, reaction-diffusion occurs at a higher temperature, and the instantaneous heat release is large, thus reaching a very high temperature and generating an Al3Pd5 phase. The elongation of the material is severely reduced when the thickness of the interfacial layer exceeds 1 μm. The optimum performance of the material was obtained at 300 °C after annealing at 300 °C for 120 s. Its tensile strength, conductivity and elongation are 140.61 MPa, 46.82%IACS and 14.89%, respectively.

Author Contributions

Software, J.G.; writing—original draft, J.G.; writing—review and editing, Z.Y., X.Y., H.X., L.P., W.Z. and X.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Technological Innovation 2025 Major Special Project of NingBo (No. 2020Z039).

Institutional Review Board Statement

Not Applicable.

Informed Consent Statement

Not Applicable.

Data Availability Statement

The raw data cannot be shared at this time as the data are also part of an ongoing study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jin, Y.H.; Wu, B.; Lu, X.T. Effect of Post-Weld Annealing on Microstructure and Growth Behavior of Copper/Aluminum Friction Stir Welded Joint. Materials 2020, 13, 4591. [Google Scholar] [CrossRef]
  2. Camacho, A.M.; Rodríguez-Prieto, Á.; Herrero, J.M.; Aragón, A.M.; Bernal, C.; Lorenzo-Martin, C.; Yanguas-Gil, Á.; Martins, P.A. An Experimental and Numerical Analysis of the Compression of Bimetallic Cylinders. Materials 2019, 12, 4094. [Google Scholar] [CrossRef] [PubMed]
  3. Chen, S.Y.; Zhang, L.; Ma, S.N. Effects of Transition Layer on Bending Resistance in Copper-Clad Aluminum Composite Casting. J. Mater. Eng. Perform. 2019, 28, 3560–3566. [Google Scholar] [CrossRef]
  4. Lechner, S.; Rœmisch, U.; Nitschke, R.; Gensch, F.; Mueller, S. Optimization of the Indirect Extrusion Process of Copper-Clad Aluminum Rods by Methods of Statistical Experimental Designs and Numerical Analyses. Front. Mater. 2021, 8, 663306. [Google Scholar] [CrossRef]
  5. Dashti, A.; Keller, C.; Vieille, B.; Guillet, A.; Bouvet, C. Experimental and Finite Element Analysis of the Tensile Behavior of Architectured Cu-Al Composite Wires. Materials 2021, 14, 6305. [Google Scholar] [CrossRef]
  6. Liu, X.H.; Liu, X.F.; Xie, J.X. Effects of annealing process on interface and mechanical properties of silver clad aluminum wires. Procedia Eng. 2012, 27, 502–511. [Google Scholar] [CrossRef]
  7. Shi, B.B.; Liu, X.H.; Xie, J.X.; Xie, M. Preparation process of silver clad aluminum bars by vertical continuous casting composite forming. Chin. J. Eng. 2019, 41, 633–645. [Google Scholar]
  8. Pang, S.L.; Chen, X.; Xu, J.S. Analysis on damage characteristics and detonation performance of solid rocket engine charge subjected to jet. Def. Technol. 2022, 18, 1552–1562. [Google Scholar] [CrossRef]
  9. Boshra, I.K.; Ahmed, A. Composite Solid Rocket Propellant Based on GAP Polyurethane Matrix with Different Binder Contents. Chin. J. Explos. Propellents 2020, 43, 362–367. [Google Scholar]
  10. Cumming, D.R.S.; Thoms, S.; Weaver, J.M.R. 3 nm NiCr wires made using electron beam lithography and PMMA resist. Microelectron. Eng. 1996, 30, 423–425. [Google Scholar] [CrossRef]
  11. Chen, K.; Ren, Q.B.; Cheng, J.M. Structural response analysis of a solid rocket motor with HTPB propellant grain under vertical storage condition. J. Northwest. Polytech. Univ. 2022, 40, 56–61. [Google Scholar] [CrossRef]
  12. Caveny, L.H.; Kuo, K.; Shackelford, B.W. Thrust and ignition transients of the Space Shuttle solid rocket motor. J. Spacecr. Rocket. 1980, 17, 489–494. [Google Scholar] [CrossRef]
  13. Iwano, K.; Hashiba, K.; Nagae, J. Reduction of tunnel blasting induced ground vibrations using advanced electronic detonators. Tunn. Undergr. Space Technol. 2020, 105, 103556. [Google Scholar] [CrossRef]
  14. Shen, Q.F. Properties and applications of reactive PdRu-Al bimetallic fuses. Rare Met. Cem. Carbides 1980, 3, 14–16. [Google Scholar]
  15. Mcalister, A.J. The Al-Pd (Aluminum-Palladium) system. Bull. Alloy. Phase Diagr. 1986, 7, 368–374. [Google Scholar] [CrossRef]
  16. Ramos, A.S.; Vieira, M.T. Intermetallic compound formation in Pd/Al multilayer thin films. Intermetallics 2012, 25, 70–74. [Google Scholar] [CrossRef]
  17. Li, M.; Li, C.; Wang, F. Thermodynamic assessment of the Al–Pd system. Intermetallics 2006, 14, 39–46. [Google Scholar] [CrossRef]
  18. Guo, J.Y. Study of PdRu5-AlMg5 composite wire diffusion layers and ignition generators. Rare Met. Mater. Eng. 1982, 4, 29–37+120–122. [Google Scholar]
  19. Howard, J.K. Kinetics of compound formation in thin film couples of Al and transition metals. Vacuum 1976, 26, 68–71. [Google Scholar] [CrossRef]
  20. Köster, U.; Ho, P.S.; Ron, M. Thermal reactions between aluminum and palladium layered films. Thin Solid Films 1980, 67, 35–44. [Google Scholar] [CrossRef]
  21. Hung, L.S.; Nastasi, M.; Gyulai, J.; Mayer, J.W. Ioninduced amorphous and crystalline phase formation in Al/Ni, Al/Pd, and Al/Pt thin films. Appl. Phys. Lett. 1983, 42, 672. [Google Scholar] [CrossRef]
  22. Peng, X.K.; Heness, G.; Yeung, W.Y. Effect of rolling temperature on interface and bond strength development of roll bonded copper/aluminum metal laminates. J. Mater. Sci. 1999, 34, 277–281. [Google Scholar] [CrossRef]
  23. Peng, X.K.; Wuhrer, R.; Heness, G.; Yeung, W.Y. Rolling strain effects on the interlaminar properties of roll bonded copper/aluminum metal laminates. J. Mater. Sci. 2000, 35, 4357–4363. [Google Scholar] [CrossRef]
  24. Zhang, H.A.; Chen, G. Fabrication of Cu/Al compound material by solid-liquid bonding method and interface bonding mechanism. Chin. J. Nonferrous Met. 2008, 18, 414–419. [Google Scholar]
  25. Lee, W.B.; Bang, K.S.; Jung, S.B. Effects of intermetallic compound on the electrical and mechanical properties of friction welded Cu/Al bimetallic joints during annealing. J. Alloys Compd. 2005, 390, 212–219. [Google Scholar] [CrossRef]
  26. Hug, E.; Bellido, N. Brittleness study of intermetallic (Cu, Al) layers in copper-clad aluminium thin wires. Mater. Sci. Eng. A 2011, 528, 7103–7106. [Google Scholar] [CrossRef]
  27. Wang, Q.N.; Liu, X.H.; Liu, X.F. Study on Annealing process and microstructure properties of cold drawn copper clad Aluminum wire. Acta Metall. Sin. 2008, 44, 30–35. [Google Scholar]
  28. Abbasi, M.; Taheri, A.K.; Salehi, M.T. Growth rate of intermetallic compounds in Al/Cu bimetal produced by cold roll welding process. J. Alloys Compd. 2001, 319, 233–241. [Google Scholar] [CrossRef]
  29. Moisy, F.; Gueydan, A.; Sauvage, X.; Kaller, C.; Guilmeau, U. Influence of intermetallic compounds on the electrical resistivity of architectured copper clad aluminum composites elaborated by a restacking drawing method. Mater. Des. 2018, 155, 366–374. [Google Scholar] [CrossRef]
Figure 1. SEM images of Pd/Al composite wire: (a) ×500, (b) ×10,000.
Figure 1. SEM images of Pd/Al composite wire: (a) ×500, (b) ×10,000.
Materials 16 01545 g001
Figure 2. SEM images of Pd/Al composite wires annealed at 200 °C for different times; (a) 5 s, (b) 60 s, (c) 7200 s, and (d) 10,800 s.
Figure 2. SEM images of Pd/Al composite wires annealed at 200 °C for different times; (a) 5 s, (b) 60 s, (c) 7200 s, and (d) 10,800 s.
Materials 16 01545 g002
Figure 3. TEM plots of Pd/Al composite wires annealed at 200 °C for different times and their SEAD plots; (a) HAADF plot at 5 s, (b) high-resolution plot of spot 1 in (a,c) FFT transform of the red region in (b,d) EDS line sweep at 5 s, (e) HAADF plot at 1080 s, (f) SAED plot of spot 4 in (e,g) SAED plot of spot 5 in (e) and EDS results for each position, and (h) EDS line sweep at 10,800 s.
Figure 3. TEM plots of Pd/Al composite wires annealed at 200 °C for different times and their SEAD plots; (a) HAADF plot at 5 s, (b) high-resolution plot of spot 1 in (a,c) FFT transform of the red region in (b,d) EDS line sweep at 5 s, (e) HAADF plot at 1080 s, (f) SAED plot of spot 4 in (e,g) SAED plot of spot 5 in (e) and EDS results for each position, and (h) EDS line sweep at 10,800 s.
Materials 16 01545 g003
Figure 4. SEM images of Pd/Al composite wires annealed at 250 °C for different times; (a) 5 s, (b) 30 s, (c) 1800 s, (d) 3600 s, and (e) 10,800 s.
Figure 4. SEM images of Pd/Al composite wires annealed at 250 °C for different times; (a) 5 s, (b) 30 s, (c) 1800 s, (d) 3600 s, and (e) 10,800 s.
Materials 16 01545 g004
Figure 5. TEM plots of Pd/Al composite wires annealed at 250 °C for different times and their SEAD plots: (a) HAADF plot at 5 s, (b) high-resolution plot of spot 1 in (a,c) FFT transform of the red region in (b,d) EDS line sweep at 5 s, (e) HAADF plot at 1080 s, (f) SAED plot of spot 2 in (e,g) SAED plot of spot 3 in (e) and EDS results for each position, and (h) EDS line sweep at 10,800 s.
Figure 5. TEM plots of Pd/Al composite wires annealed at 250 °C for different times and their SEAD plots: (a) HAADF plot at 5 s, (b) high-resolution plot of spot 1 in (a,c) FFT transform of the red region in (b,d) EDS line sweep at 5 s, (e) HAADF plot at 1080 s, (f) SAED plot of spot 2 in (e,g) SAED plot of spot 3 in (e) and EDS results for each position, and (h) EDS line sweep at 10,800 s.
Materials 16 01545 g005
Figure 6. SEM images of Pd/Al composite wires annealed at 300 °C for different times; (a) 5 s, (b) 60 s, (c) 900 s, (d) 3600 s, (e) 7200 s, and (f) 10,800 s.
Figure 6. SEM images of Pd/Al composite wires annealed at 300 °C for different times; (a) 5 s, (b) 60 s, (c) 900 s, (d) 3600 s, (e) 7200 s, and (f) 10,800 s.
Materials 16 01545 g006
Figure 7. TEM plots of Pd/Al composite wires annealed at 300 °C for different times and their SEAD plots. (a) HAADF plot at 5 s, (b) high-resolution plot of spot 1 in (a,c) FFT transform of the red region in (b,d) EDS line sweep at 5 s, (e) HAADF plot at 1080 s and EDS results for each position, and (f) SAED plot of spot 3 in (e,g) EDS line sweep at 10,800 s.
Figure 7. TEM plots of Pd/Al composite wires annealed at 300 °C for different times and their SEAD plots. (a) HAADF plot at 5 s, (b) high-resolution plot of spot 1 in (a,c) FFT transform of the red region in (b,d) EDS line sweep at 5 s, (e) HAADF plot at 1080 s and EDS results for each position, and (f) SAED plot of spot 3 in (e,g) EDS line sweep at 10,800 s.
Materials 16 01545 g007
Figure 8. SEM images of Pd/Al composite wires annealed at 350 °C for different times; (a) 5 s, (b) 60, (c) 120 s, (d) 300 s, (e) 3600 s, and (f) 10,800 s.
Figure 8. SEM images of Pd/Al composite wires annealed at 350 °C for different times; (a) 5 s, (b) 60, (c) 120 s, (d) 300 s, (e) 3600 s, and (f) 10,800 s.
Materials 16 01545 g008
Figure 9. TEM plots of Pd/Al composite wires annealed at 350 °C for different times and their SEAD plots. (a) HAADF plot at 5 s, (b) high-resolution plot of spot 1 in (a,c) FFT transform of the red region in (b,d) EDS line sweep at 5 s, (e) HAADF plot at 1080 s, (f) SAED plot of spot 2 in (e,g) SAED plot of spot 3 in (e) and EDS results for each position, and (h) EDS line sweep at 10,800 s.
Figure 9. TEM plots of Pd/Al composite wires annealed at 350 °C for different times and their SEAD plots. (a) HAADF plot at 5 s, (b) high-resolution plot of spot 1 in (a,c) FFT transform of the red region in (b,d) EDS line sweep at 5 s, (e) HAADF plot at 1080 s, (f) SAED plot of spot 2 in (e,g) SAED plot of spot 3 in (e) and EDS results for each position, and (h) EDS line sweep at 10,800 s.
Materials 16 01545 g009
Figure 10. Properties of Pd/Al composite wires at different temperatures with annealing time: (a) tensile strength, (b) elongation, and (c) electrical conductivity.
Figure 10. Properties of Pd/Al composite wires at different temperatures with annealing time: (a) tensile strength, (b) elongation, and (c) electrical conductivity.
Materials 16 01545 g010
Table 1. Interfacial intermetallic compounds generated under different annealing conditions for Pd/Al composite wires.
Table 1. Interfacial intermetallic compounds generated under different annealing conditions for Pd/Al composite wires.
T(°C)Time(s)
510,800
200Al3Pd2Al3Pd2, Al21Pd8
250Al3Pd2Al3Pd2, Al21Pd8
300Al3Pd2Al3Pd2
350Al3Pd5Al3Pd2, AlPd
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gui, J.; Yang, Z.; Yin, X.; Xie, H.; Peng, L.; Zhang, W.; Mi, X. Effect of Annealing on the Interface and Properties of Pd/Al Composite Wires. Materials 2023, 16, 1545. https://doi.org/10.3390/ma16041545

AMA Style

Gui J, Yang Z, Yin X, Xie H, Peng L, Zhang W, Mi X. Effect of Annealing on the Interface and Properties of Pd/Al Composite Wires. Materials. 2023; 16(4):1545. https://doi.org/10.3390/ma16041545

Chicago/Turabian Style

Gui, Jiabin, Zhen Yang, Xiangqian Yin, Haofeng Xie, Lijun Peng, Wenjing Zhang, and Xujun Mi. 2023. "Effect of Annealing on the Interface and Properties of Pd/Al Composite Wires" Materials 16, no. 4: 1545. https://doi.org/10.3390/ma16041545

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop