Next Article in Journal
Effects of the Primary NbC Elimination on the SSCC Resistance of a HSLA Steel for Oil Country Tubular Goods
Previous Article in Journal
Implants Survival Rate in Regenerated Sites with Innovative Graft Biomaterials: 1 Year Follow-Up
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Texture Evolution and Nanohardness in Cu-Nb Composite Wires

1
Shenyang National Laboratory for Materials Science, Chongqing University, Chongqing 400044, China
2
International Joint Laboratory for Light Alloys (Ministry of Education), Department of Materials Science and Engineering, Chongqing University, Chongqing 400044, China
*
Author to whom correspondence should be addressed.
Materials 2021, 14(18), 5294; https://doi.org/10.3390/ma14185294
Submission received: 10 August 2021 / Revised: 8 September 2021 / Accepted: 10 September 2021 / Published: 14 September 2021

Abstract

:
Multifilamentary microcomposite copper-niobium (Cu-Nb) wires were fabricated by a series of accumulative drawing and bonding steps (ADB). The texture of the Cu matrix in these wires was studied using electron backscattered diffraction (EBSD) and transmission electron microscopy (TEM). Dynamic recrystallization during cold drawing caused a weakening of the <111> texture in the micron-scale Cu matrix at high values of true strain. A sharp <111> texture was observed in the nano-scale Cu matrix due to the suppression of dynamic recrystallization. The grain size was reduced by the higher level of dynamic recrystallization at high strains. The relation between the nanoindentation behavior of the different Cu matrix and the grain sizes, Cu-Nb interface, and texture was established.

1. Introduction

Over the past two decades, Cu-Nb microcomposites have been widely used in defense, aerospace, and magnetic applications for their excellent combination of mechanical properties, high conductivity, and thermal stability [1,2,3,4]. Several techniques have been used to fabricate these materials, including accumulative drawing and bonding (ADB) [5,6], melt and deform [7,8], accumulative roll bonding (ARB) [9,10,11], and magnetron sputtering [12,13]. Compared with other techniques, the ADB process enables the production of Cu-Nb microcomposite wires more than 100 m in length [14]. During this process, a multi-scale Cu matrix with grain sizes ranging from a few microns to a few nanometers is formed [15]. Metals experiencing axisymmetric deformation after ADB will develop a fiber texture and only one direction is needed to fully represent the preferred orientation [16,17]. The large number of embedded Nb fibers affects the deformation and texture development of the Cu matrix [18]. A variety of constraints, including the Cu-Nb interface area and the matrix dimensions, will alter the deformation compared to non-composite FCC materials [19]. It is well established that the texture exerts a strong influence on the mechanical properties and electrical conductivity of such materials. For example, wires with a <111> fiber texture will exhibit higher stiffness than wire with a random orientation [20,21,22]. Thus, studies on texture evolution are required to better understand the process–structure relationships.
The microstructure and texture evolution of Cu-Nb microcomposites have been widely investigated [6,8,19,23,24,25,26]. Using X-ray diffraction (XRD) and neutron diffraction, researchers [6,23,25] observed that a sharp <111> fiber texture with a weak <100> texture for the Cu matrix and a distinct <110> texture for Nb developed in these composite wires. The intensities of these two fiber texture components in the Cu matrix are intricately affected by intermediate annealing and deformation strain, whereas the texture of Nb is basically unchanged [23,24]. Popova et al. [25] found that both <111> and <100> texture components are weakened at higher cold drawing strains due to dynamic recovery and recrystallization in Cu-Nb composite wires. However, it is impossible to characterize the texture at different scales in the Cu matrix by XRD. Several studies [19,26] utilized TEM to investigate the crystallographic relationship between the Cu matrix and Nb fibers. Deng et al. [19] reported that a comparatively parallel relationship of <111> Cu and <110> Nb with a few degrees of deviation formed in Cu-Nb wires at a strain of 24.8. Nevertheless, the texture obtained by TEM did not support this result statistically. Recently, an EBSD investigation of as-deformed Cu-Nb wires [15,16] at different scales demonstrated different proportions of <111> and <100> texture [15] within regions of the Cu matrix. However, these studies mainly focused on the micron-scale texture of the Cu matrix rather than the nanoscale. It seems that a complete understanding of texture evolution for Cu-Nb microcomposite wires during deformation requires further systematic research.
ADB Cu-Nb microcomposite wires are well known as a bi-phase material with a multi-scale microstructure. Conventional mechanical testing methods are not well suited to studying the mechanical properties of Cu channels at different scales. Nanoindentation testing has been used extensively for multi-scale and dual-phase materials to analyze the mechanical properties with high spatial resolution [27,28,29,30,31,32]. Using nanoindentation, Thilly et al. [30] found that Cu-Nb interfaces have a strong blocking effect on dislocation slip and attributed the high strength of Cu-Nb wires to a size effect and interface strengthening. Nanoindentation was applied along the rolling (RD) and transverse (TD) directions of an ARB Cu-Nb nanolaminate to study the effect of anisotropy [32]. Unfortunately, there are few nanoindentation studies focused directly on the Cu matrix in composite wires at different strains.
In this paper, the evolution of each layer of the Cu channel was characterized in detail at different levels of strain. The principles of texture formation at the different scales of the Cu matrix are discussed. Nanoindentation testing was applied to quantify the hardness values for different Cu channels. The influential effects of the Cu matrix, grain sizes, Cu-Nb interface, and texture on the mechanical properties of the microcomposite wires were established.

2. Materials and Methods

High-purity niobium (99.9%) and oxygen-free high-conductivity (OFHC) copper were utilized to fabricate Cu-Nb composite wires with a fixed number of Nb fibers using an ADB process. The ADB process results in a multi-scale Cu matrix containing continuous parallel Nb fibers with a maximum number of up to 854 (N = 854, determined by the number of drawing steps) and with diameters varying in the range of 10 nm to 500 μm. Four different types of specimens were prepared after each pass of the first four ADB process, with N = 85, 852, 853, and 854. These samples were designated by the number of Nb fibers. The ADB process can be seen in detail elsewhere [3].
EBSD specimens were cut from the ADB wires, and electron polished in a solution of 83.3% H3PO4/16.7% H2O at a voltage of 2 V (<0.01 A) and 20 °C. The EBSD characterization was carried out in a scanning electron microscope (TESCAN MIRA 3, Tescan Corporation, Brno, Czech Republic) with step sizes from 0.01 to 1.5 μm. TEM samples were prepared by a low-temperature ion-thinning technique and examined in an FEI Tecnai F20 electron microscope (Thermo Fisher Scientific, Waltham, MA, USA) at a voltage of 200 kV.
Nanoindentation tests were performed on each channel of copper in different samples using a Hysitron Triboindenter (Hysitron Inc., Minneapolis, MN, USA) and a Berkovich indenter. A peak load of 6 mN and loading time of 5 s were applied with loading and unloading rates of 12 mN/min. The positions of indentation were selected using an optical microscope (OM) equipped in Triboindenter. Atomic force microscopy (AFM, Hysitron Inc., Minneapolis, MN, USA) was performed to identify the surface profile of the post-test indents. More than 30 tests were carried out in each test area to get accurate nanohardness values. The indentation regions were observed via SEM to determine the area fraction of each phase. All the indentation tests were carried out in the longitudinal (drawing) direction.

3. Results

3.1. Microstructure and Texture

The cross-sectional SEM microstructure of the Cu-Nb composite wire containing 854 of Nb fibers is shown in Figure 1. The 854 Nb fibers were embedded in a Cu matrix with five different spacings of Cun: from the largest outer-most jacket (n = 0) to the finest nanoscale Cu (n = 4). Similarly, the 853 Nb fibers were embedded in a Cu matrix with four spacings of Cun (from Cu0 to Cu3). Similar results were also observed in the 852 and 85 samples. Based on the distance between Nb fibers, the scale of the copper matrix could be divided into three categories: micro-scale copper (M-Cu, larger than 1 μm), sub-micron-scale copper (S-Cu, 100 nm–1 μm), and nanoscale copper (N-Cu, less than 100 nm). The dimensional parameters and characteristics of the Cu-Nb composite wires at different ADB stages are displayed in Table 1. A linear intercept method was used to measure the dimension of the Cui and Nb fibers based on the SEM and TEM images (only dCu3, dCu4, and dNb in 854).
EBSD maps (at different scales) of the Cu matrix in the four Cu-Nb wires are shown in Figure 2. A strong <111> and <100> fiber texture developed in the Cu matrix during the drawing deformation, a result consistent with those in the literature [23,24]. The total proportion of <100> and <111> texture increased gradually with the strain. There were grains with other orientations in the 852 Cu1, yet it was nearly invisible in the other Cu channels (Cu2, Cu3 and Cu4). A more extensive substructure was observed in the 854 Cu2 sample compared to 854 Cu3. Notably, a higher fraction of substructure developed in the <111> grains than in the <100> grains.
Figure 3a displays the normalized statistical results of the <111> and <100> texture components. A comparison between different samples with the same level of Cu at M-Cu (Cu0 and Cu2) shows that the <111> texture weakened generally with increasing true strain, whereas the <111> texture in Cu1 showed a contrary tendency. In each sample, the degree of deformation due to the ADB process increased from Cu1 to Cu4, and the <111> texture increased at first (from Cu0 to Cu1) and then decreased (from Cu1 to Cu3). In 854 Cu3, the proportion of <111> texture decreased to less than 10%. In nanoscale Cu4, however, more than 60% of the grains were highly <111> textured. That is, the intensity of the <111> texture was enhanced at strains larger than 19.1, showing a contrary tendency compared to previous work by Popova et al. [25]. Thus, it is necessary to investigate the relation of drawing strain and the matrix size on the development of texture.
The grain sizes for the Cu channels in different samples are shown in Figure 3b. Severe dynamic recrystallization resulted in a weaker <111> texture and grain refinement. As the Cu channel size was reduced to a few hundred nanometers, the Cu-Nb interfaces began to play an important role. Single Cu grain filled the space between niobium fibers. This reduction in grain size decreased with the increase in dynamic recrystallization.
Figure 4 shows the bright field images of copper regions in 854. Large dislocation cells were observed in M-Cu, as shown in Figure 4a, indicating that dynamic recrystallization occurred in the Cu matrix at high deformation. The Nb fibers and Cu channels are marked in Figure 4b,c. There were significant differences in the interface where the Cu matrix width was more than 100 nm (S-Cu) or less than 100 nm (N-Cu), as shown in Figure 4b. Insignificant contrast induced by tangled dislocations was observed at the Cu-Nb interface where the Cu matrix width was less than 100 nm. The straight white dashed line marks the Cu-Nb interface in an area where a large number of dislocation tangles was observed.

3.2. Nanoindentation Tests

Given that the dCu3 and dCu4 in 853 and 854 samples were less than 3 μm, it was difficult for the indenter to cover only Cu3 or Cu4 zone. Therefore, only the nanohardness of Cui (i < 3) was directly tested in the present work. The nanohardness of Cu3 and Cu4 was calculated using a modified rule of mixtures (ROM) [30,33], taking the multi-scale nature of the composites into account:
H com = X Nb H Nb + X Cu 3 H Cu 3 + X Cu 4 H Cu 4
where Hcom is the nanohardness of the tested regions and Xα is the volume fraction of the Nb or Cui (as seen in Table 2); the reference data for Nb crystals at micron(Hμ-Nb) and nanometer scales (Hn-Nb) were 2.4 and 4.2 GPa, respectively [30]. Therefore, for the 853 sample:
H cu 3 = H Cu 3 + Nb X μ Nb H μ Nb X Cu 3
As for the 854 sample, where dCu3, dCu4 and dNb are in the nano-scale range, the HCu3 and HCu4 could be calculated:
H cu 4   =   H Cu 4 + Nb X n Nb H n Nb X Cu 4
H cu 3   =   H Cu 3 + Cu 4 + Nb X Cu 4 + Nb H Cu 4 + Nb X Cu 3
The nanohardness of Cu regions in the different Cu-Nb wires is illustrated in Figure 5. The grain size decreased drastically with the Cu channel dimension from the Cu0 region of the 85 sample to the Cu2 region of the 854 sample. However, no significant change of nanohardness was observed in any of the examined Cui (i < 3) regions. The average hardness was measured to be 1.72 ± 0.14 GPa, with a maximum value of 2.06 GPa and a minimum of 1.42 GPa. These results are lower than the experimental results from Thilly [30], but closer to the data in the literature [34]. The nanohardness of Cu3 in the 853 sample was about 2.65 GPa, and the hardness of Cu3 and Cu4 in 854 were 2.88 and 3.10 GPa, respectively. Compared with Cui (i < 3), the ultra-high hardness of Cu3 in 853 and Cu3/Cu4 in 854 indicate that there were other strengthening mechanisms in play.
The depth ratio after unloading and at peak load (hf/hmax) is usually used to interpret the indentation behavior. Typical load/displacement curves of the examined samples are shown in Figure 6. The hf, hmax, and other parameters are derived from these curves. Pharr et al. [35,36] concluded that when the hf/hmax ratio is less than 0.8, the pile-up can be ignored. That is, the nanohardness value is reliable when hf/hmax < 0.8. Figure 7 displays AFM images and surface profiles at four typical indentation regions. Obviously, only in the nanocomposite regions (Cu3+Cu4+Nb/Cu4+Nb regions in 854) was the hf/hmax less than 0.8. Little pile-up was found in these regions (Figure 7f,h). In contrast, extended pile-up around the indent was observed in the regions with hf/hmax > 0.8 (Figure 7b,d).

4. Discussion

4.1. Texture Analysis

The formation of <100> texture in micro-scale Cu during drawing is related to dynamic recovery and recrystallization, driven by deformation energy stored in the form of dislocations [37]. At low strain levels, dislocation pileups at grain boundaries or entangled dislocation substructures are developed. As the true strain increases, dislocations accumulate at low-angle sub-grain boundaries. High-angle grain boundaries are formed as the misorientation gradually increases. This mechanism is fully illustrated in the dynamic recrystallization process of M-Cu. Lee et al. [38] studied the effect of temperature on fiber texture, suggesting that recrystallization leads to the formation of <100> texture. Yang et al. [39] considered the effect of cold drawing on the microstructure evolution of Cu wires and found that dynamic recrystallization took place at strains greater than 1.91. In the present work, the smallest strain in the Cu channel was 4.7, much larger than 1.91. Dynamic recrystallization at this strain level led to a decrease in <111> texture and increase in <100> texture in M-Cu and S-Cu. As depicted in Figure 8a, the grains of Cu3 in 854 exhibited a recrystallized morphology with a strong <100> texture. The less-developed substructure developed in the <100> grains of 854 Cu2 also supports this interpretation.
As shown in Figure 3a, the proportion of <100> texture in the Cu3 channel in the 853 sample was more than 60%. At higher deformation levels, the recrystallized <100> grains rotated to the <111> orientation [40], resulting in a stronger <111> texture. Meanwhile, the occurrence of dynamic recovery and recrystallization enhanced the <100> texture. After the last ADB step, the Cu3 in 853 became Cu4 in 854. Grain rotation was competed with dynamic recovery and recrystallization during the high-strain deformation.
Figure 8b presents the transverse sectional microstructure of Cu4+Nb in 854. Compared with the equiaxed structure in Cu3 in 854, most grains of Cu4 underwent severe deformation. Instead of Cu grain boundaries, a high density of Cu-Nb interfaces were created in the Cu4+Nb region. Such interfaces often play an important role in determining the properties of composite materials [1,19,41,42] and are especially important in composites with nanoscale microstructures, since a large number of defects are present [43,44]. Misfit dislocations and vacancies can act as obstacles for dislocation movement and are ideal sinks for dislocation annihilation [44]. When the Cu matrix size is reduced to submicron scale, the effect of the interface becomes more potent. Deng et al. [19] found a sharp increase in interface density within the nano-scale microstructure of a Cu-Nb composite. Figure 3a shows that the 854 sample with both nanoscales and sub-micron scales of Cu4 exhibited a sharp <111> texture. In Figure 4b, the contrast induced by dislocation tangles in the Cu matrix was observed near the Cu-Nb interface where the Cu channel width was greater than 150 nm. A considerable amount of N-Cu in Cu4 was created due to the high strain of Cu4 (finest Cu channel) and the limitation of the Cu-Nb interfaces. Insignificant dislocation tangles were observed in the Cu channels when the width was close to or less than 100 nm (Figure 4c). Tangled dislocations can be eliminated by mutual annihilation or interface diffusion [45]. The deformation mechanism of the nanoscale Cu-Nb wire was confined to a layer slip that involved the movement of single dislocation loops on parallel planes between Nb fibers [46,47], which indicates that there was basically no dislocation entanglement in the nanoscale Cu matrix. Thus, dynamic recrystallization in these nanoscale Cu channels is more difficult due to the low driving force, and the deformation-induced <111> texture of the nanoscale Cu matrix was preserved.

4.2. Nanoindentation Behavior

Thilly et al. [30] reported that the averaged hardness of the largest Cu channels (dCu > 10 μm) is 2.17 GPa. This overestimated hardness is attributed to the pile-up around the indent. The pile-up can underestimate the calculated contact area between indenter and materials, leading to overestimation of the hardness values. In the present study, the hf/hmax was close to 0.88 in Cui (i < 3) channels, resulting in an underestimation of the calculated contact area and a positive deviation of the actual hardness. Although the pile-up phenomena could overestimate the hardness when the hf/hmax > 0.8, the change trend of hardness in different regions was not affected.
Microstructure dimension d, such as grain size and phase spacing, has a significant effect on the nanohardness [30,34,48]. Generally, the hardness is not dependent on size when d > 10 μm, whereas significant hardening occurs as d is reduced from 10 to 1 μm [30]. The dCu2 in 853 and 854 composite wires was less than 10 μm. However, compared with other Cu channels with larger microstructure size, the hardness increase seemed unusually high in both samples. Li et al. [34] investigated the hardness of nano–micro-structured bulk copper by nanoindentation, and the results showed that the hardness values of coarse-grained Cu and nanocrystalline Cu were about 1.1 GPa and 2.1 GPa, respectively. The minimum grain size of Cui (i < 3) in the tested areas in the 854 composite wire was about 0.5 μm due to dynamic recrystallization; thus, the hardness was close to that of coarse-grained Cu, inferring that nanohardness is similar regardless of the scale of the Cu matrix.
The calculated hardness values of Cu3 in 853 and Cu3/Cu4 in 854 were higher than HCui (i < 3). However, since the hf/hmax decreased with the size of the microstructure dimension, an overestimated hardness due to pile-up cannot explain the ultra-high hardness in these areas. For 853, the grain size in Cu3 was less than 300 nm, leading to a positive deviation of the calculated hardness since the grain refinement strengthening. When the dNb was about 3 μm, Nb grains were nano-grains, resulting in an underestimation of Hμ-Nb calculated based on Equation (2). This is another possible explanation for the ultra-high hardness of Cu3 in 853. For 854, the dα (α = Cu3, Cu4, Nb) was smaller than 1 μm, and there was little pile-up in the Cu3+Cu4+Nb/Cu4+Nb region. However, a large number of Cu-Nb interfaces existed with the potential for a large influence on the obvious hardness values [19,47]. Previous research has demonstrated that microhardness increases with the interface area density. Therefore, the unusually high nanohardness observed in this study was quite possibly a result of the interface density.
In addition, crystallographic orientation has a great effect on the hardness response in nanoindentation tests [27,28,29]. Wang et al. studied the mechanical properties of single-crystal Cu with different crystallographic orientations through nanoindentation, revealing that the nanohardness of (100)-, (110)-, and (111)-oriented single crystalline Cu are 1.33, 1.48, and 1.20 GPa, respectively [28]. Roa et al. studied the relationship between γ-(fcc) austenite grain orientation and nanoindentation hardness in metastable stainless steel, and found that (100) and (110) grains have lower hardness than (111) grains [27]. In the present study, the Cu4 channel in 854 had strong <111> fiber texture compared to other Cu channels, which could also have resulted in the abnormally high level of hardness.

5. Conclusions

The texture and nanohardness of the different Cu channels in ADB Cu-Nb microcomposite wires were investigated. The main conclusions are as follows:
(1)
Two fiber texture components, i.e., <111> and <100>, were found in the Cu matrix. In the micro-scale Cu matrix, the <111> texture weakened at higher levels of strain; for instance, the proportion of the <111> texture of Cu3 in 854 was less than 10%. In submicron and nanoscale Cu channels, a sharp <111> texture was observed due to the suppression of dynamic recrystallization at the interface. The grain size was affected by dynamic recrystallization and decreased as the degree of deformation increased. The grain size was reduced by the higher level of dynamic recrystallization at high strains.
(2)
The nanohardness of Cui (i < 3) in different Cu-Nb composite wires was similar, even though their grain sizes were orders of magnitude different. The slightly overestimated hardness value was caused by the pile-up near the indentation. There was no obvious size effect when dCui > 3 μm.
(3)
Cu3 and Cu4 in 853/854 exhibited ultra-high nanohardness. For Cu3 in the 853 sample, this result was mainly ascribed to the size effect and the underestimation of HNb in the calculation process. The size effect, high density of Cu-Nb interfaces, and texture effects were suggested for the unusually high nanohardness of Cu3 and Cu4 in the 854 Cu-Nb composite wires.

Author Contributions

Conceptualization, S.X., Y.L. and X.Y.; methodology, S.X. and Y.L.; software, S.X. and L.W.; validation, S.X., X.Y. and L.W.; formal analysis, S.X., Y.L. and L.W.; investigation, S.X.; resources, X.Y.; data curation, S.X.; writing—original draft preparation, S.X. and Y.L.; writing—review and editing, X.Y.; visualization, S.X.; supervision, X.Y.; project administration, X.Y.; funding acquisition, X.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant No. 51571046 and No. 51421001), and Project No. 2020CDJDCL001, supported by the Fundamental Research Funds for the Central Universities.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors appreciate the support from the 111 Project (B16007) by the Ministry of Education and the State Administration of Foreign Experts Affairs of China. The authors also gratefully acknowledge the access to the facilities of the Electron Microscopy Center and Analytical and Testing Center in Chongqing University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chen, Y.; Li, N.; Hoagland, R.G.; Liu, X.Y.; Baldwin, J.K.; Beyerlein, I.J.; Cheng, J.Y.; Mara, N.A. Effects of three-dimensional Cu/Nb interfaces on strengthening and shear banding in nanoscale metallic multilayers. Acta Mater. 2020, 199, 593–601. [Google Scholar] [CrossRef]
  2. Spencer, K.; Lecouturier, F.; Thilly, L.; Embury, J.D. Established and emerging materials for use as high-field magnet conductors. Adv. Eng. Mater. 2004, 6, 290–297. [Google Scholar] [CrossRef]
  3. Gu, T.; Medy, J.R.; Volpi, F.; Castelnau, O.; Forest, S.; Herve-Luanco, E.; Lecouturier, F.; Proudhon, H.; Renault, P.-O.; Thilly, L. Multiscale modeling of the anisotropic electrical conductivity of architectured and nanostructured Cu-Nb composite wires and experimental comparison. Acta Mater. 2017, 141, 131–141. [Google Scholar] [CrossRef] [Green Version]
  4. Deng, L.; Han, K.; Wang, B.; Yang, X.; Liu, Q. Thermal stability of Cu–Nb microcomposite wires. Acta Mater. 2015, 101, 181–188. [Google Scholar] [CrossRef]
  5. Sun, Q.Q.; Jiang, F.; Deng, L.; Xiao, H.X.; Li, L.; Liang, M.; Peng, T. Uniaxial fatigue behavior of Cu-Nb micro-composite conductor, Part II: Modeling and simulation. Int. J. Fatigue 2016, 91, 286–292. [Google Scholar] [CrossRef] [Green Version]
  6. Wang, P.F.; Liang, M.; Xu, X.Y.; Li, J.S.; Zhang, P.X. Thermal Stability of Cu-18 vol%Nb Composites. Rare Met. Mater. Eng. 2020, 49, 1171–1176. [Google Scholar]
  7. Verhoeven, J.D.; Downing, H.L.; Chumbley, L.S.; Gibson, E.D. The resistivity and microstructure of heavily drawn Cu–Nb alloys. J. Appl. Phys. 1989, 65, 1293–1301. [Google Scholar] [CrossRef] [Green Version]
  8. Leprince-Wang, Y.; Han, K.; Huang, Y.; Yu-Zhang, K. Microstructure in Cu–Nb microcomposites. Mater. Sci. Eng. A 2003, 351, 214–223. [Google Scholar] [CrossRef]
  9. Liu, Z.; Snel, J.; Boll, T.; Wang, J.Y.; Monclús, M.A.; Molina-Aldareguía, J.M.; LLorca, J. High temperature strength retention of Cu/Nb nanolaminates through dynamic strain ageing. Mater. Sci. Eng. A 2021, 799, 140117. [Google Scholar] [CrossRef]
  10. Gao, R.; Jin, M.M.; Han, F.; Wang, B.M.; Wang, X.P.; Fang, Q.F.; Dong, Y.H.; Sun, C.; Shao, L.; Li, M.D.; et al. Superconducting Cu/Nb nanolaminate by coded accumulative roll bonding and its helium damage characteristics. Acta Mater. 2020, 197, 212–223. [Google Scholar] [CrossRef]
  11. Ding, C.G.; Xu, J.; Li, X.W.; Shan, D.B.; Guo, B.; Terence, G. Langdon. Microstructural Evolution and Mechanical Behavior of Cu/Nb Multilayer Composites Processed by Accumulative Roll Bonding. Adv. Eng. Mater. 2019, 22, 1900702. [Google Scholar] [CrossRef]
  12. Yang, W.F.; Gong, M.Y.; Yao, J.H.; Wang, J.W.; Zheng, S.J.; Ma, X.L. Hardening induced by dislocation core spreading at disordered interface in Cu/Nb multilayers. Scr. Mater. 2021, 200, 113917. [Google Scholar] [CrossRef]
  13. Wu, K.; Zhang, J.Y.; Wang, Y.Q.; Li, S.J.; Liu, G.; Sun, J. Size- and strain-dependent buckle dimensions of nanostructured Cu/Nb multilayers on polyimide substrates. Mater. Lett. 2017, 206, 44–47. [Google Scholar] [CrossRef]
  14. Embury, J.D.; Han, K. Conductor materials for high field magnets. Curr. Opin. Solid State Mater. Sci. 1998, 3, 304–308. [Google Scholar] [CrossRef]
  15. Gu, T.; Medye, J.-R.; Klosekf, V.; Castelnaua, O.; Forestb, S.; Hervé-Luancoc, E.; Lecouturier–Dupouy, F.; Proudhonb, H.; Renaulte, P.-O.; Thilly, L.; et al. Multiscale modeling of the elasto-plastic behavior of architectured and nanostructured Cu-Nb composite wires and comparison with neutron diffraction experiments. Int. J. Plast. 2019, 122, 1–30. [Google Scholar] [CrossRef] [Green Version]
  16. Deng, L.; Yang, X.; Han, K.; Lu, Y.; Liang, M.; Liu, Q. Microstructure and texture evolution of Cu–Nb composite wires. Mater. Charact. 2013, 81, 124–133. [Google Scholar] [CrossRef]
  17. Vidal, V.; Thilly, L.; Lecouturier, F.; Renault, P.O. Effects of size and geometry on the plasticity of high-strength copper/tantalum nanofilamentary conductors obtained by severe plastic deformation. Acta Mater. 2006, 54, 1063–1075. [Google Scholar] [CrossRef]
  18. Vidal, V.; Thilly, L.; Van petegem, S.; Stuhr, U.; Lecouturier, F.; Renault, P.O.; Van swygenhoven, H. Plasticity of nanostructured Cu–Nb-based wires: Strengthening mechanisms revealed by in situ deformation under neutrons. Scr. Mater. 2009, 60, 171–174. [Google Scholar] [CrossRef]
  19. Deng, L.P.; Liu, Z.F.; Wang, B.S.; Han, K.; Xiang, H.L. Effects of interface area density and solid solution on the micro-hardness of Cu-Nb microcomposite wires. Mater. Charact. 2019, 150, 62–66. [Google Scholar] [CrossRef]
  20. Min, B.; Lim, S.-S.; Jung, S.-J.; Kim, G.; Lee, B.-H.; Won, S.O.; Kim, S.K.; Rhyee, J.-S.; Kim, J.-S.; Baek, S.-H. Texture-induced reduction in electrical resistivity of p-type (Bi, Sb)2Te3 by a hot extrusion. J. Alloys Compd. 2018, 764, 261–266. [Google Scholar] [CrossRef]
  21. Renk, O.; Ghosh, P.; Sabat, R.K.; Eckert, J.; Pippan, R. The role of crystallographic texture on mechanically induced grain boundary migration. Acta Mater. 2020, 200, 404–416. [Google Scholar] [CrossRef]
  22. Gollapudi, S.; Rai, N.; Kushwaha, R.; Sabat, R.K. Crystallographic texture influences on the thermal stability of nanocrys-talline materials. J. Mater. Sci. 2021, 56, 11154–11163. [Google Scholar] [CrossRef]
  23. Carpenter, J.S.; Liu, X.; Darbal, A.; Nuhfer, N.T.; McCabe, R.J.; Vogel, S.C.; LeDonne, J.E.; Rollett, A.D.; Barmak, K.; Beyerlein, I.J.; et al. A comparison of texture results obtained using precession electron diffraction and neutron diffraction methods at diminishing length scales in ordered bimetallic nanolamellar composites. Scr. Mater. 2012, 67, 336–339. [Google Scholar] [CrossRef]
  24. Gu, T.; Castelnau, O.; Forest, S.; Hervé-Luanco, E.; Lecouturier, F.; Proudhon, H.; Thilly, L. Multiscale modeling of the elastic behavior of architectured and nanostructured Cu–Nb composite wires. Int. J. Solids Struct. 2017, 121, 148–162. [Google Scholar] [CrossRef]
  25. Popova, E.N.; Popov, V.V.; Romanov, E.P.; Hlebova, N.E.; Shikov, A.K. Effect of deformation and annealing on texture parameters of composite Cu–Nb wire. Scr. Mater. 2004, 51, 727–731. [Google Scholar] [CrossRef]
  26. Sun, I.H.; Mary, A.H. Microstructural stability of Cu–Nb microcomposite wires fabricated by the bundling and drawing process. Mater. Sci. Eng. A 2000, 281, 189–197. [Google Scholar]
  27. Roa, J.J.; Fargas, G.; Mateo, A.; Jiménez-Piqué, E. Dependence of nanoindentation hardness with crystallographic orienta-tion of austenite grains in metastable stainless steels. Mater. Sci. Eng. A 2015, 645, 188–195. [Google Scholar] [CrossRef]
  28. Wang, Z.; Zhang, J.; Hassan, H.; Zhang, J.; Yan, Y.; Hartmaier, A.; Sun, T. Coupled effect of crystallographic orientation and indenter geometry on nanoindentation of single crystalline copper. Int. J. Mech. Sci. 2018, 148, 531–539. [Google Scholar] [CrossRef]
  29. Chen, S.; Miyahara, Y.; Nomoto, A. Crystallographic orientation dependence of nanoindentation hardness in austenitic phase of stainless steel. Philos. Mag. Lett. 2019, 98, 473–485. [Google Scholar] [CrossRef]
  30. Thilly, L.; Lecouturier, F.; von Stebut, J. Size-induced enhanced mechanical properties of nanocomposite copper/niobium wires: Nanoindentation study. Acta Mater. 2002, 50, 5049–5065. [Google Scholar] [CrossRef]
  31. Zhang, L.; Liang, M.; Feng, Z.; Zhang, L.; Cao, W.; Wu, G. Nanoindentation characterization of strengthening mechanism in a high strength ferrite/martensite steel. Mater. Sci. Eng. A 2019, 745, 144–148. [Google Scholar] [CrossRef]
  32. Sahay, R.; Budiman, A.S.; Harito, C.; Gunawan, F.E.; Navarro, E.; Escoubas, S.; Cornelius, T.W.; Aziz, I.; Lee, P.S.; Thomas, O. Berkovich nanoindentation study of 16 nm Cu/Nb ARB nanolaminate: Effect of anisotropy on the surface pileup. MRS Adv. 2021, 6, 495–499. [Google Scholar] [CrossRef]
  33. Thilly, L.; Véron, M.; Ludwig, O.; Lecouturier, F. Deformation mechanism in high strength Cu-Nb nanocomposites. Mater. Sci. Eng. A 2001, 309, 510–513. [Google Scholar] [CrossRef]
  34. Li, W.L.; Tao, N.R.; Lu, K. Fabrication of a gradient nano-micro-structured surface layer on bulk copper by means of a sur-face mechanical grinding treatment. Scr. Mater. 2008, 59, 546–549. [Google Scholar] [CrossRef]
  35. Bolshakov, A.; Oliver, W.; Pharr, G. Influences of stress on the measurement of mechanical properties using nanoindenta-tion: Part II. Finite element simulations. J. Mater. Res. 1996, 11, 760–768. [Google Scholar] [CrossRef]
  36. Pharr, G. Measurement of mechanical properties by ultra-low load indentation. Mater. Sci. Eng. A 1998, 253, 151–159. [Google Scholar] [CrossRef]
  37. Humphreys, F.J.; Rohrer, G.S.; Rollett, A.D. Recrystallization and Related Annealing Phenomena, 3rd ed.; Elsevier: Oxford, UK, 2017; pp. 169–213. [Google Scholar]
  38. Lee, D.N.; Chung, Y.H.; Shin, M.C. Preferred orientation in extruded aluminum-alloy rod. Scr. Mater. 1983, 17, 339–342. [Google Scholar]
  39. Yang, F.; Dong, L.M.; Cai, L.; Wang, L.F.; Xie, Z.H.; Fang, F. Effect of cold drawing strain on the microstructure, mechanical properties and electrical conductivity of low-oxygen copper wires. Mater. Sci. Eng. A 2021, 818, 141348. [Google Scholar] [CrossRef]
  40. Chen, J.A.; Yan, W.; Li, W.; Miao, J.A.; Fan, X.H. Texture evolution and its simulation of cold drawing copper wires produced by continuous casting. Trans. Nonferrous Met. Soc. China 2011, 21, 152–158. [Google Scholar] [CrossRef]
  41. Zheng, S.J.; Beyerlein, I.J.; Carpenter, J.S.; Kang, K.W.; Wang, J.; Han, W.Z.; Mara, N.A. High-strength and thermally stable bulk nanolayered composites due to twin-induced interfaces. Nat. Commun. 2013, 4, 1696. [Google Scholar] [CrossRef] [Green Version]
  42. Liang, Y.; Yang, X.; Gong, M.; Liu, G.; Liu, Q.; Wang, J. Slip transmission for dislocations across incoherent twin boundary. Scr. Mater. 2019, 166, 39–43. [Google Scholar] [CrossRef]
  43. Demkowicz, M.J.; Thilly, L. Structure, shear resistance and interaction with point defects of interfaces in Cu–Nb nanocom-posites synthesized by severe plastic deformation. Acta Mater. 2011, 59, 7744–7756. [Google Scholar] [CrossRef] [Green Version]
  44. Beyerlein, I.J.; Demkowicz, M.J.; Misra, A.; Uberuaga, B.P. Defect-interface interactions. Prog. Mater. Sci. 2015, 74, 125–210. [Google Scholar] [CrossRef] [Green Version]
  45. Zhu, X.Y.; Luo, J.T.; Chen, G.; Zeng, F.; Pan, F. Size dependence of creep behavior in nanoscale Cu/Co multilayer thin films. J. Alloys Compd. 2010, 506, 434–440. [Google Scholar] [CrossRef]
  46. Thilly, L.; Véron, M.; Ludwig, O.; Lecouturier, F.; Peyrade, J.P.; Askénazy, S. High-strength materials:in-situ investigations of dislocation behaviour in Cu-Nb multifilamentary nanostructured composites. Philos. Mag. A 2002, 82, 925–942. [Google Scholar] [CrossRef]
  47. Wang, J.; Misra, A. An overview of interface-dominated deformation mechanisms in metallic multilayers. Curr. Opin. Solid State Mater. Sci. 2011, 15, 20–28. [Google Scholar] [CrossRef]
  48. Olugbade, T.O.; Lu, J. Literature review on the mechanical properties of materials after surface mechanical attrition treat-ment (SMAT). Nano Mater. Sci. 2020, 2, 3–31. [Google Scholar] [CrossRef]
Figure 1. SEM micrograph showing the multi-scale microstructure of N = 854 Cu-Nb microcomposite wires. In the maximum magnification view, the Nb fibers exhibit light contrast embedded in a dark-contrast Cu matrix (removed by etch).
Figure 1. SEM micrograph showing the multi-scale microstructure of N = 854 Cu-Nb microcomposite wires. In the maximum magnification view, the Nb fibers exhibit light contrast embedded in a dark-contrast Cu matrix (removed by etch).
Materials 14 05294 g001
Figure 2. EBSD maps of copper matrix with different scales in different samples. (a-1) Cu0 in 851; (a-2) Cu1 in 851; (b-1) Cu0 in 852; (b-2) Cu1 in 852; (b-3) Cu2 in 852; (c-1) Cu0 in 853; (c-2) Cu1 in 853; (c-3) Cu2 in 853; (c-4) Cu3 in 853; (d-1) Cu0 in 854; (d-2) Cu1 in 854; (d-3) Cu2 in 854; (d-4) Cu3 and Cu4 in 854.
Figure 2. EBSD maps of copper matrix with different scales in different samples. (a-1) Cu0 in 851; (a-2) Cu1 in 851; (b-1) Cu0 in 852; (b-2) Cu1 in 852; (b-3) Cu2 in 852; (c-1) Cu0 in 853; (c-2) Cu1 in 853; (c-3) Cu2 in 853; (c-4) Cu3 in 853; (d-1) Cu0 in 854; (d-2) Cu1 in 854; (d-3) Cu2 in 854; (d-4) Cu3 and Cu4 in 854.
Materials 14 05294 g002
Figure 3. (a) Normalized area fraction of <111> and <100> texture in copper matrix, (b) grain size of copper matrix in different samples.
Figure 3. (a) Normalized area fraction of <111> and <100> texture in copper matrix, (b) grain size of copper matrix in different samples.
Materials 14 05294 g003
Figure 4. Bright field TEM images of 854 sample (a) dislocation cells in M-Cu, (b) longitudinal sections showing Nb fibers surrounded by S-Cu and N-Cu, and (c) longitudinal sections showing clean N-Cu and Nb fibers.
Figure 4. Bright field TEM images of 854 sample (a) dislocation cells in M-Cu, (b) longitudinal sections showing Nb fibers surrounded by S-Cu and N-Cu, and (c) longitudinal sections showing clean N-Cu and Nb fibers.
Materials 14 05294 g004
Figure 5. Nanohardness data of copper channels in different Cu-Nb wires.
Figure 5. Nanohardness data of copper channels in different Cu-Nb wires.
Materials 14 05294 g005
Figure 6. Typical load/displacement curves for different test areas of Cu-Nb composite wires.
Figure 6. Typical load/displacement curves for different test areas of Cu-Nb composite wires.
Materials 14 05294 g006
Figure 7. AFM images and corresponding surface profiles of indentations in different positions of Cu-Nb wires: (a,b) Cu1 channel in 852; (c,d) Cu3+Nb region in 853; (e,f) Cu3+Cu4+Nb region in 854; (g,h) Cu4+Nb region in 854.
Figure 7. AFM images and corresponding surface profiles of indentations in different positions of Cu-Nb wires: (a,b) Cu1 channel in 852; (c,d) Cu3+Nb region in 853; (e,f) Cu3+Cu4+Nb region in 854; (g,h) Cu4+Nb region in 854.
Materials 14 05294 g007
Figure 8. Transverse sectional HAADF-STEM of 854 Cu-Nb microcomposite wire, (a) Cu3+Cu4+Nb region, (b) Cu4+Nb region, (c) EDS elemental mapping, and (d) EDS spectrum of (b).
Figure 8. Transverse sectional HAADF-STEM of 854 Cu-Nb microcomposite wire, (a) Cu3+Cu4+Nb region, (b) Cu4+Nb region, (c) EDS elemental mapping, and (d) EDS spectrum of (b).
Materials 14 05294 g008
Table 1. Dimensional characteristics of the selection of Cu-Nb wires.
Table 1. Dimensional characteristics of the selection of Cu-Nb wires.
SampleNd (mm)%CudCu1 (μm)dCu2 (μm)dCu3/μmdCu4 (nm)dNb (μm)η
851856.7462149 ± 15428 ± 209.6
8528525.57 *7363 ± 1414 ± 441 ± 814.4
8538535.64 *7963 ± 199 ± 22 ± 13 ± 119.1
8548545.068460 ± 165 ± 2462 ± 170 (nm)97 ± 60119 ± 90 (nm)24.8
Volume Fractions
853XNb = 0.21 XCu0 = 0.222 XCu1 = 0.225 XCu2 = 0.124 XCu3 = 0.219
854XNb = 0.16 XCu0 = 0.238 XCu1 = 0.169 XCu2 = 0.172 XCu3 = 0.095 XCu4 = 0.167
Notes: N, the Nb filament number; d, the total diameter of the Cu-Nb wire (* hexagon); %Cu, the total volume fraction of Cu; dCui, the measured width of the Cui channel; dNb, the width of the Nb fibers; η, the total logarithmic true strain. (η = ln(A0/A), where A is the final cross-sectional area of the sample after deformation and A0 is the cross-sectional area of the wire before drawing). Xα, the volume fraction of α.
Table 2. Volume fraction and nanohardness of the composite regions in 853 and 854 Cu-Nb wires.
Table 2. Volume fraction and nanohardness of the composite regions in 853 and 854 Cu-Nb wires.
Tested RegionsXCu3XCu4XNbH (GPa)
853 Cu3+Nb0.510.492.53 ± 0.19
854 Cu3+Cu4+Nb0.2250.3950.383.47 ± 0.16
854 Cu4+Nb0.510.493.64 ± 0.21
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xiang, S.; Yang, X.; Liang, Y.; Wang, L. Texture Evolution and Nanohardness in Cu-Nb Composite Wires. Materials 2021, 14, 5294. https://doi.org/10.3390/ma14185294

AMA Style

Xiang S, Yang X, Liang Y, Wang L. Texture Evolution and Nanohardness in Cu-Nb Composite Wires. Materials. 2021; 14(18):5294. https://doi.org/10.3390/ma14185294

Chicago/Turabian Style

Xiang, Shihua, Xiaofang Yang, Yanxiang Liang, and Lu Wang. 2021. "Texture Evolution and Nanohardness in Cu-Nb Composite Wires" Materials 14, no. 18: 5294. https://doi.org/10.3390/ma14185294

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