Next Article in Journal
Fracture Process Analysis and Acoustic Emission Response of Cemented Tailings Backfill with Different Sizes under Uniaxial Compression
Next Article in Special Issue
Impact of Titanium Addition on Microstructure, Corrosion Resistance, and Hardness of As-Cast Al+6%Li Alloy
Previous Article in Journal
Oscillating Laser Post-Processing of NiCrCoFeCBSi/WC Thermally Sprayed Coatings
Previous Article in Special Issue
Predicting Elastic Constants of Refractory Complex Concentrated Alloys Using Machine Learning Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Cold Working on the Properties and Microstructure of Cu-3.5 wt% Ti Alloy

1
School of Materials Science and Engineering, Northeastern University, Shenyang 110819, China
2
State Key Laboratory of Nonferrous Metals and Processes, GRIMAT Group Co., Ltd., Beijing 101407, China
3
GRIMAT Engineering Institute Co., Ltd., Beijing 101407, China
4
General Research Institute for Nonferrous Metals, Beijing 100088, China
5
China Nonferrous Metals Mining Group (Tianjin) Co., Ltd., Tianjin 300112, China
*
Authors to whom correspondence should be addressed.
Materials 2022, 15(22), 8042; https://doi.org/10.3390/ma15228042
Submission received: 10 October 2022 / Revised: 29 October 2022 / Accepted: 31 October 2022 / Published: 14 November 2022

Abstract

:
Cu-Ti alloys were strengthened by β’-Cu4Ti metastable precipitation during aging. With the extension of the aging time, the β’-Cu4Ti metastable phase transformed into the equilibrium β-Cu4Ti phase. The Cu-3.5 wt% Ti(Cu-4.6 at% Ti) alloys with different processing were aged at different temperatures for various times after solution treatment at 880 °C for 1 h. The electrical conductivity of samples under different heat treatments had shown an upward trend as time increased during aging, but the hardness reached the peak value and then decreased. The hardness and electrical conductivity of the samples with 70% deformation after aging are higher tha n the samples without deformation. Deformation after aging would cause the metastable phase to dissolve into a matrix. The best combination value of conductivity and hardness is 13.88% IACS and 340.78 Hv, and the optimal heat treatment is 500 °C for 2 h + 70% deformation + 450 °C for 2 h.

1. Introduction

Compared with other alloys, copper alloys have high electrical conductivity even after alloying or heat treatment to improve mechanical properties. Therefore, copper alloys, such as wires and connectors, are excellent electrical materials. There are many ways to strengthen high-conductivity alloys, such as alloying, precipitation and grain refinement, but these strengthening methods lead to a decrease in conductivity [1,2,3]. For industrial applications of copper alloys, a balance between economy, mechanical properties and electrical conductivity should be considered, so the most effective method is precipitation hardening [4,5,6]. The well-known precipitation-strengthening copper alloys are Cu-Be [7,8], Cu-Ni-Si [9,10] and Cu-Ti [11,12,13,14] alloys. Cu-Ti alloy has high strength of 1 GPa and relatively high electrical conductivity compared with other copper alloys, so it is now used as connectors or conductive spring materials and has a wider range of applications as the increasing demand for high strength and high electrical conductivity [15,16,17,18]. Cu-Ti alloy undergoes spinodal decomposition during aging, forming the metastable phase β’-Cu4Ti (tetragonal structure with I4/m space group), which transforms into the stable phase β-Cu4Ti (orthorhombic structure with Pmmn space group) with prolonging aging time. During aging, the metastable phase β’-Cu4Ti precipitates within grains, and it transforms to stable β-Cu4Ti with a layered structure at the grain boundaries with the increase of aging time. The former is called continuous precipitation, and the latter is called discontinuous precipitation. Compared to other copper alloys, Cu-Ti alloy needs a longer aging time to obtain the proper combination of strength and electrical conductivity and has a lower electrical conductivity [19,20]. Therefore, it is necessary to introduce deformation processing, but the influence of deformation processing on the properties and microstructure of Cu-Ti alloy is still unclear. This paper focuses on the influence of deformation on the properties and microstructure of Cu-Ti binary alloy.

2. Experimental Procedure

Cu-3.5Ti (wt.%) ingots were obtained by smelting and casting pure copper and titanium in an intermediate frequency furnace with dimensions of 300 × 150 × 20 mm. The chemical compositions of the alloys were determined through inductively coupled plasma atomic emission spectrometry (ICP-AES), as listed in Table 1. Following the removal of the surface defects, the ingot was hot rolled at 850 °C from 20 to 5 mm. Thereafter, the hot-rolled samples of the Cu-3.5Ti alloy were solid solution-treated at 880 °C for 60 min to obtain a supersaturated solid solution. To confirm the optimal aging treatment, the alloy samples were aged at different temperatures at different times. The solid solution samples were isothermally aged from 400 to 525 °C for 1 h. The most suitable initial aging temperature and time are determined. After the initial aging treatment, the sample is deformed by 70% to 1.5 mm and then aged at different temperatures for a long time. The processing process diagram is shown in Figure 1. Subsequently, to remove surface defects, the ingot was milled to 1.2 mm. Moreover, the hardness of these samples, which were cut to the same size (10 × 10 × 1.2 mm) by wire cutting, was tested using a VH1150 Macro Vickers Hardness Tester with a 5 kg load for 15 s. Further, the conductivity tests of samples were conducted using an eddy current conductivity instrument Sigma 2008. Each sample was tested at least 5 times, and the values were averaged. The microstructure of samples with different aging times was characterized using a JEM-F200 field emission transmission electron microscopy (TEM) system. These samples were processed into 3 mm disks with 80 μm thickness followed by electrolytic double jet polishing from −40 to −50 °C using a solution of CH3OH:HNO3 = 70%:30%. Further, the X-ray diffraction (XRD) data were obtained using a SmartLab XRD instrument.

3. Results

Figure 2a shows the variation curves of the electrical conductivity and Vickers hardness of the Cu-3.5Ti alloy aged for 1 h from 400 to 525 °C. The electrical conductivity of Cu-3.5Ti alloys increased with the temperature rising during aging. The hardness of the alloy reached a peak value of 295.91 HV at 500 °C. Thereafter, the hardness of the alloy decreased with an increase in temperature over 500 °C, thereby indicating the occurrence of over-aging. As shown in Figure 2b, the optimum aging time at 500 °C is 2 h. Subsequent experiments will be conducted on samples with optimal aging.
The bright-field image and corresponding SAED pattern of Cu-3.5Ti at the peak aging state are shown in Figure 3. Long rod-like phases with a width of 26.4 nm are observed along the [001]Cu direction. From the SAED pattern, the diffraction patterns are confirmed as those of the β’-Cu4Ti phase in Figure 3b. The orientation relationship between the precipitation phase and the matrix can be obtained. The orientation relationships between the Cu matrix and β’-Cu4Ti phase were [001]Cu//[001]β’ and (100)Cu//(310)β’ at the peak aging time.
Figure 4 shows the bright-field image and corresponding SAED pattern of Cu-3.5Ti aged at 500 °C for 8 h. The discontinuous precipitates with lamellar structures are also found. From the SAED pattern, the diffraction patterns are confirmed as those of the β-Cu4Ti phase in Figure 4b. It indicates the formation of equilibrium discontinuous precipitates at the over-aging state. The orientation relationships between β-Cu4Ti and the Cu matrix were [110]Cu//[102]β and (111)Cu//(010)β.
Cold deformation was carried out on the samples in peak aging state. With the increase in deformation, the samples change from uniform deformation to non-uniform deformation, as shown in Figure 5. In the uniform deformation stage, the number of twin boundaries obviously increases, and in the non-uniform deformation stage, slip bands perpendicular to the rolling direction appear.
The samples with 70% cold deformation were aged at different temperatures and times, and their hardness and electrical conductivity were measured and shown in Figure 6. Taking the balance between conductivity and hardness into consideration, the best aging treatment of samples with 70% deformation is 450 °C for 2 h. Compared with the samples aged directly without deformation, the hardness and conductivity of samples with 70% deformation are both improved obviously, and the aging temperature is reduced correspondingly.

4. Discussion

The samples with different degrees of deformation after peak aging were tested by XRD. It was found that the peak corresponding to β’-Cu4Ti metastable precipitation gradually disappeared with the increased degrees of deformation, as shown in Figure 7. It indicates that for the deformed samples, the metastable β’-Cu4Ti phase had been dissolved, which is consistent with the significant decrease in the conductivity of the deformed sample from 12.62% IACS to 5.45% IACS.
The bright-field image of 70% deformed samples shows a large number of twin boundaries in the matrix, and the width of the β’-Cu4Ti phase decreases obviously, as shown in Figure 8. The twin diffraction spots can be seen from the SEAD pattern, but the spots of the metastable β’-Cu4Ti phase had disappeared, which also proves that the deformation after aging will make the β’-Cu4Ti phase dissolve into the matrix. Figure 9 shows the TEM image and element distribution maps of samples aging at 450 °C for 2 h after 70% deformation. The formation of the metastable β’-Cu4Ti phase can be clearly observed in Figure 9c, and the width of the precipitated phase is 16.3 nm, smaller and denser than the direct aging sample after solid solution treatment. Therefore, the property of the samples with 70% deformation is more excellent.
It can be concluded that the twin boundaries and slip bands generated by the deformation act as a heterogeneous nucleation site and accelerate the precipitation of β’-Cu4Ti, leading to an increase in hardness and conductivity at the same time. In addition, when combined with the change in hardness and conductivity according to the aging curve shown in Figure 6, the deformation before aging was able to effectively reduce the temperature to reach the highest hardness during aging. In this study, it has been found that aging with deformation processing is very effective for Cu-Ti alloys to improve both hardness and conductivity. It can be interpreted that Cu-Ti alloys are precipitated through spinodal decomposition. Spinodal decomposition occurs during aging, in which the solid solution is separated into a rich portion and a lean portion. This chemical composition forms a stress field in a microscopic range and inhibits the movement of dislocations or shear bands, which increases strength. Meanwhile, these defects act as heterogeneous nucleation sites. The precipitation of the β’-Cu4Ti phase is accelerated, and the β’-Cu4Ti phase is finer. It is interpreted as increasing the strength and conductivity within the alloy. As such, the aging with deformation processing in the Cu-Ti alloy reduces the precipitation temperature and increases the highest hardness. It can be an effective method to simultaneously improve strength and conductivity.

5. Conclusions

In this study, in order to examine the effect of the deformation of Cu-3.5wt% Ti alloy on aging, it was processed at 70% deformation after solid solution treatment and aged at 500 °C for 2 h. The change in hardness and electrical conductivity with aging was analyzed, and the reason was examined through microstructure analysis The following conclusions were drawn:
  • Cu-3.5Ti alloys were strengthened by the β’-Cu4Ti metastable phase during aging, with the extension of the aging time, the metastable phase transformed into the equilibrium β-Cu4Ti phase.
  • Deformation after aging would cause the metastable phase to dissolve into the matrix, and the twins and slip bands generated during deformation would accelerate the precipitation of the metastable β’-Cu4Ti phase in aging treatment, reduce the precipitation of β’-Cu4Ti phase temperature of the precipitation phase, and make the β’-Cu4Ti phase finer.
  • The optimal heat treatment is 500 °C for 2 h + 70% deformation + 450 °C for 2 h, the corresponding conductivity and hardness are 13.88% IACS and 340.78 Hv.

Author Contributions

L.H.: Investigation, writing—original draft, writing—review& editing. L.P.: Conceptualization, writing—review & editing. G.Z.: Conceptualization, supervision. G.H.: Supervision, investigation. H.X.: Conceptualization, supervision. X.M.: Conceptualization, supervision, validation, resources. W.Z.: Supervision, validation. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Technological Innovation 2025 & Major Special Project of NingBo (No. 2021Z084 and 2020Z039).

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Laughlin, D.E.; Cahn, J.W. Spinodal decomposition in age hardening copper-titanium alloys. Acta Metall. 1975, 23, 329–339. [Google Scholar] [CrossRef]
  2. Datta, A.; Soffa, W.A. The structure and properties of age hardened Cu-Ti alloys. Acta Metall. 1976, 24, 987–1001. [Google Scholar] [CrossRef]
  3. Soffa, W.A.; Laughlin, D.E. High-strength age hardening copper–titanium alloys: Redivivus. Prog. Mater. Sci. 2001, 49, 347–366. [Google Scholar] [CrossRef]
  4. Knights, R.; Wilkes, P. The precipitation of titanium in copper and copper-nickel base alloys. Acta Metall. 1973, 21, 1503–1514. [Google Scholar] [CrossRef]
  5. Sinning, H.R.; Haasen, P. The fatigue behaviour of spinodally decomposed Cu-4at% Ti single crystals. Scr. Metall. 1981, 15, 85–90. [Google Scholar] [CrossRef]
  6. Rouxel, B.; Cayron, C.; Bornand, J.; Sanders, P.; Logé, R.E. Micro-addition of Fe in highly alloyed Cu-Ti alloys to improve both formability and strength. Mater. Des. 2022, 213, 110340. [Google Scholar] [CrossRef]
  7. Lomakin, I.; Castillo-Rodríguez, M.; Sauvage, X. Microstructure, mechanical properties and aging behaviour of nanocrystalline copper–beryllium alloy. Mater. Sci. Eng. A 2019, 744, 206–214. [Google Scholar] [CrossRef]
  8. Hariram, M.; Theerath, D.; Chakravarthy, P.; Kumar, R.A. Influence of Cold work on aging response of C17200-Beryllium Copper Alloy C17200. Mater. Today Proc. 2017, 4, 11188–11193. [Google Scholar] [CrossRef]
  9. Liu, F.; Li, J.; Peng, L.; Huang, G.; Xie, H.; Ma, J.; Mi, X. Simultaneously enhanced hardness and electrical conductivity in a Cu-Ni-Si alloy by addition of Cobalt. J. Alloys Compd. 2021, 862, 158667. [Google Scholar] [CrossRef]
  10. Lei, Q.; Li, Z.; Wang, M.P.; Zhang, L.; Xiao, Z.; Jia, Y.L. The evolution of microstructure in Cu-8.0Ni-1.8Si-0.15Mg alloy during aging. Mater. Sci. Eng. A 2010, 527, 6728–6733. [Google Scholar] [CrossRef]
  11. Wang, X.; Xiao, Z.; Qiu, W.-t.; Li, Z.; Liu, F. The evolution of microstructure and properties of a Cu–Ti–Cr–Mg–Si alloy with high strength during the multi-stage thermomechanical treatment. Mater. Sci. Eng. A 2020, 803, 140510. [Google Scholar] [CrossRef]
  12. Nagarjuna, S.; Sharma, K.K.; Sudhakar, I.; Sarma, D.S. Age hardening studies in a Cu–4.5Ti–0.5Co alloy. Mater. Sci. Eng. A 2001, 313, 251–260. [Google Scholar] [CrossRef]
  13. Monzen, R.; Hosoda, T.; Takagawa, Y.; Watanabe, C. Bend formability and strength of Cu-Be-Co alloys. J. Mater. Sci. 2011, 46, 4284–4289. [Google Scholar] [CrossRef] [Green Version]
  14. Liu, W.; Li, J.; Chen, X.; Ji, M.; Xiao, X.; Wang, H.; Yang, B. Effect of vanadium on the microstructure and kinetics of discontinuous precipitation in Cu–3.2Ti–0.2Fe alloy. J. Mater. Res. Technol. 2021, 14, 121–136. [Google Scholar] [CrossRef]
  15. Semboshi, S.; Takasugi, T. Fabrication of high-strength and high-conductivity Cu–Ti alloy wire by aging in a hydrogen atmosphere. J. Alloys Compd. 2013, 580, S397–S400. [Google Scholar] [CrossRef]
  16. Huang, L.; Cui, Z.; Meng, X.; Zhang, X.; Zhang, X.; Song, X.; Tang, N.; Xiao, Z.; Lei, Q.; Li, Z. Effects of microelements on the microstructure evolution and properties of ultrahigh strength Cu–Ti alloys. Mater. Sci. Eng. A 2021, 823, 141581. [Google Scholar] [CrossRef]
  17. Semboshi, S.; Kaneno, Y.; Takasugi, T.; Masahashi, N. Suppression of discontinuous precipitation in Cu-Ti alloys by aging in a hydrogen atmosphere. Metall. Mater. Trans. A 2020, 51, 3704–3712. [Google Scholar] [CrossRef]
  18. Semboshi, S.; Kaneno, Y.; Takasugi, T.; Masahashi, N. High strength and high electrical conductivity Cu-Ti alloy wires fabricated by aging and severe drawing. Metall. Mater. Trans. A 2018, 49, 4956–4965. [Google Scholar] [CrossRef]
  19. Si, L.; Zhou, L.; Zhu, X.; Sanhua, L.; Leinuo, S.; Qiyi, D. Microstructure and property of Cu–2.7Ti–0.15Mg–0.1Ce–0.1Zr alloy treated with a combined aging process. Mater. Sci. Eng. A 2016, 650, 345–353. [Google Scholar] [CrossRef]
  20. Fukamachi, K. Kinetics and thermodynamics associated with age hardening of Cu–4at%Ti alloy. Mater. Sci. Eng. A 2020, 788, 139411. [Google Scholar] [CrossRef]
Figure 1. Schematic diagrams of thermo-mechanical treatment processes.
Figure 1. Schematic diagrams of thermo-mechanical treatment processes.
Materials 15 08042 g001
Figure 2. Hardness and electrical conductivity curves of the Cu-3.5Ti alloy aged at (a) different temperatures for 1 h and (b) 500 °C for different times.
Figure 2. Hardness and electrical conductivity curves of the Cu-3.5Ti alloy aged at (a) different temperatures for 1 h and (b) 500 °C for different times.
Materials 15 08042 g002
Figure 3. TEM and SAED images of the Cu-3.5Ti alloy at the peak aging state. (a) Bright-field image; (b) SAED pattern.
Figure 3. TEM and SAED images of the Cu-3.5Ti alloy at the peak aging state. (a) Bright-field image; (b) SAED pattern.
Materials 15 08042 g003
Figure 4. TEM images of the Cu-3.5Ti alloy aging at 500 °C for 8 h. (a) Bright-field image; (b) SAED pattern.
Figure 4. TEM images of the Cu-3.5Ti alloy aging at 500 °C for 8 h. (a) Bright-field image; (b) SAED pattern.
Materials 15 08042 g004
Figure 5. Optical photographs with different degrees of deformation (a) 5%; (b) 10%; (c) 25%; (d) 70%.
Figure 5. Optical photographs with different degrees of deformation (a) 5%; (b) 10%; (c) 25%; (d) 70%.
Materials 15 08042 g005
Figure 6. (a) Hardness and (b) electrical conductivity curves of the Cu-3.5Ti alloy aged at 400 °C to 550 °C with 70% deformation.
Figure 6. (a) Hardness and (b) electrical conductivity curves of the Cu-3.5Ti alloy aged at 400 °C to 550 °C with 70% deformation.
Materials 15 08042 g006
Figure 7. XRD patterns of samples with different degrees of deformation.
Figure 7. XRD patterns of samples with different degrees of deformation.
Materials 15 08042 g007
Figure 8. TEM and SAED images of the peak aging condition of Cu-3.5Ti alloy with 70% deformation (a) Bright-field image; (b) SAED pattern.
Figure 8. TEM and SAED images of the peak aging condition of Cu-3.5Ti alloy with 70% deformation (a) Bright-field image; (b) SAED pattern.
Materials 15 08042 g008
Figure 9. TEM image and corresponding element distribution (a) Bright-field image; (b) distribution of Cu; (c) distribution of Ti.
Figure 9. TEM image and corresponding element distribution (a) Bright-field image; (b) distribution of Cu; (c) distribution of Ti.
Materials 15 08042 g009
Table 1. Chemical composition of the Cu-3.5Ti alloy (wt.%).
Table 1. Chemical composition of the Cu-3.5Ti alloy (wt.%).
Nominal CompositionTiCu
Cu-3.5Ti3.58Bal.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Huang, L.; Peng, L.; Mi, X.; Zhao, G.; Huang, G.; Xie, H.; Zhang, W. Effect of Cold Working on the Properties and Microstructure of Cu-3.5 wt% Ti Alloy. Materials 2022, 15, 8042. https://doi.org/10.3390/ma15228042

AMA Style

Huang L, Peng L, Mi X, Zhao G, Huang G, Xie H, Zhang W. Effect of Cold Working on the Properties and Microstructure of Cu-3.5 wt% Ti Alloy. Materials. 2022; 15(22):8042. https://doi.org/10.3390/ma15228042

Chicago/Turabian Style

Huang, Lue, Lijun Peng, Xujun Mi, Gang Zhao, Guojie Huang, Haofeng Xie, and Wenjing Zhang. 2022. "Effect of Cold Working on the Properties and Microstructure of Cu-3.5 wt% Ti Alloy" Materials 15, no. 22: 8042. https://doi.org/10.3390/ma15228042

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