Next Article in Journal
Scanning Electron Microscopy Investigation of Surface Acoustic Wave Propagation in a 41° YX-Cut of a LiNbO3 Crystal/Si Layered Structure
Next Article in Special Issue
Synthesis and Characterization of Antibacterial Carbopol/ZnO Hybrid Nanoparticles Gel
Previous Article in Journal
Enhancement of Planar Orientation of Reactive Mesogen Molecules for Optical Retarder Film by Anisotropic Surface Plasma Treatment
Previous Article in Special Issue
Crystalline Silicon Spalling as a Direct Application of Temperature Effect on Semiconductors’ Indentation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Analysis of Microstructure and Mechanical Properties of Bi-Modal Nanoparticle-Reinforced Cu-Matrix

1
Mechanical Department, Faculty of Technology and Education, Suez University, Suez 43519, Egypt
2
Powder Technology Division, Manufacturing Technology Department, Central Metallurgical R & D Institute, 1 Elfelezat St. Eltebeen, Cairo 11421, Egypt
3
Mechanical Engineering Department, Egyptian Academy for Engineering & Advanced Technology, Affiliated to Ministry of Military Production, Cairo 3056, Egypt
4
Nanotechnology Lab El Nozha, Electronic Research Institute (E.R.I.), Cairo 12622, Egypt
*
Author to whom correspondence should be addressed.
Crystals 2021, 11(9), 1081; https://doi.org/10.3390/cryst11091081
Submission received: 25 July 2021 / Revised: 26 August 2021 / Accepted: 2 September 2021 / Published: 6 September 2021

Abstract

:
Bi-modal particles are used as reinforcements for Cu-matrix. Nano TiC and/or Al2O3 were mechanically mixed with Cu particles for 24 h. The Cu-TiC/Al2O3 composites were successfully produced using spark plasma sintering (SPS). To investigate the effect of TiC and Al2O3 nanoparticles on the microstructure and mechanical properties of Cu-TiC/Al2O3 nanocomposites, they were added, whether individually or combined, to the copper (Cu) matrix at 3, 6, and 9 wt.%. The results showed that titanium carbide was homogeneously distributed in the copper matrix, whereas alumina nanoparticles showed some agglomeration at Cu grain boundaries. The crystallite size exhibited a clear reduction as a reaction to the increase of the reinforcement ratio. Furthermore, increasing the TiC and Al2O3 nanoparticle content in the Cu-TiC/Al2O3 composites reduced the relative density from 95% for Cu-1.5 wt.% TiC and 1.5 wt.% Al2O3 to 89% for Cu-4.5 wt.% TiC and 4.5 wt.% Al2O3. Cu-9 wt.% TiC achieved a maximum compressive strength of 851.99 N/mm2. Hardness values increased with increasing ceramic content.

1. Introduction

Copper strengthening is a current priority due to the pressing need to use it in various applications requiring a balance of properties [1,2,3]. Metal-matrix composites are most promising in achieving balanced mechanical properties between nano and microstructure materials [4,5,6,7,8]. Copper is used in many industries owing to its low cost, ease of manufacturing, and good corrosion resistance [9]. The main drawbacks of pure copper are its substantial low strength, high coefficient of thermal expansion (CTE), and generally poor mechanical properties [10]. One effective way to overcome these limitations is to reinforce copper with ceramic particles to obtain composites with superior properties. The effectiveness of dispersed particles in matrix strengthening depends primarily on particle characteristics: size, distribution, spacing, thermodynamic stability, and low solubility and diffusivity of its constituent elements in the matrix. Among ceramic particles, alumina nanoparticles have shown outstanding mechanical properties even at high temperatures, as well as low production costs [11,12]. In addition, TiC is an attractive candidate for metallic matrices such as copper (Cu), iron (Fe), aluminum (Al), titanium (Ti), and nickel (Ni) because of its high hardness, high melting point, and abrasion resistance with good electrical conductivity [12,13,14]. Due to the aforementioned factors, Cu reinforced with (TiC-Al2O3) composites led to a more viable material.
Numerous techniques have been used to fabricate reinforced copper matrix composites (CMCs), including molecular-level mixing (MLM) [15], in situ metallurgy [12,13], flake powder metallurgy [16,17], high-energy ball milling (HEBM) [7,18,19,20], friction stir processing [21,22,23,24,25,26], high-pressure torsion [26], and rolling [27,28,29,30]. Although these techniques enhanced the mechanical properties of processed composites, they resulted in an inhomogeneous distribution of particle reinforcements within the matrix. Additionally, they have the potential to cause morphological and structural damage, as demonstrated through carbon nanotubes (CNTs) within a copper matrix [31].
The spark plasma sintering (SPS) method, developed recently, is a new technique for synthesizing metal matrix composites. The SPS technique has piqued researchers’ interest due to its advantages of sintering at relatively low temperatures, higher heating speeds, shorter processing times, and the absence of pre-compression as in conventional sintering. Thus, the SPS technique enables the fabrication of nanostructured composites without the high grain growth rate associated with traditional sintering methods. As a result, SPS composites exhibit exceptional mechanical properties at room temperature, even at elevated temperatures [32].
To the authors’ knowledge, few papers discuss the solid-state spark plasma sintered Cu-Al2O3 [33] and Cu-TiC [34,35,36], respectively. However, no information on the synthesis and mechanical investigation of hybrid Cu-Al2O3-TiC through mechanical alloying and SPS techniques have been released. Thus, this work fabricated three separate nanocomposites of Cu-TiC, Cu-Al2O3, and hybrid Cu-TiC-Al2O3, using mechanical alloying and SPS processing. The influence of the TiC and Al2O3 nanoparticles content on the microstructure and mechanical properties of the prepared nanocomposites was also investigated.

2. Materials and Methods

2.1. Materials

Copper (Cu) powder with 99.9% purity (supplied by AlphaChemical, MA, USA) with an average particle size of 10 μm was used as a metal matrix. Alumina (Al2O3) nanopowder with 99.7% purity (supplied by Alpha Chemicals, MA, USA) with an average particle size of 50 nm and titanium carbide (TiC) nanopowder with 99.7% purity (supplied by Inframat Advanced Materials, L.L.C., CT, USA) with an average size of 100 nm were used. Both TiC and Al2O3 were used as individual/hybrid reinforcement. The Cu powder was mixed with 3, 6, and 9 wt.% of individual/hybrid reinforcement of TiC and Al2O3 using a ball milling technique for 24 h. The powders were mixed in a stainless-steel vial and protected from oxidation using highly pure argon gas using a 25:1 ball to powder ratio (BPR), 110 rpm, and a ball diameter of 5 mm. Stearic acid (1.5 wt.%) was used as a process controlling agent (PCA). Figure 1 and Table 1 show the composition of fabricated samples.

2.2. Spark Plasma Sintering (SPS)

The sintering process was performed using a spark plasma sintering technique (DR. SINTER LAB Model: SPS-1030, Syntex, Osaka, Japan). In all experiments, the powder was loaded into a graphite die with an inner diameter of 15 mm with graphite foil and enclitic by 0.5 mm thick graphite cover to prevent the friction of the sample with the die during the compaction process and to minimize heat loss. Before sintering, the SPS chamber was evacuated to a pressure below 5 Pa. The samples were heated from room temperature up to 950 °C by pulsed D.C. current using the heating rate of 20 °C/min. The samples were then held at the maximum temperature for 45 min under a uniaxial pressure of 30 MPa applied since the first minute of heating. This processing route was used to fabricate the Cu-TiC/Al2O3 nanocomposites.

2.3. Mechanical Properties

The hardness was measured along the polished surface of the specimen using a Vickers hardness tester (HMV-2T Model SHIMADZU, Kyoto, Japan). The test was carried out under 100 g load for 15 s dwell time.
The microhardness values were evaluated for an average of twelve readings on the surface of each sample. The compression test for the investigated specimens was carried out using a universal testing machine. In the compression test, three samples were investigated, and average results were obtained. The dimensions of the specimens for compression tests were 6 mm in diameter and 15 mm in length. The applied crosshead speed was 0.05 mm/s, and the test was performed at room temperature.

3. Results

3.1. XRD Analysis

Figure 2 shows the XRD patterns of the prepared ten samples, pure Cu and 3, 6, 9 wt.% TiC/Al2O3 (individual and hybrid) nanocomposites. Only peaks corresponding to Cu, TiC, and Al2O3 appeared, whereas pattern-like Cu was observed in the case of 3 wt.% Cu/TiC/Al2O3 samples; this may be due to the lower percentage of both TiC and Al2O3 that are below the limits of the XRD device. This may be attributed to the controlled milling and sintering process in an argon atmosphere which shows that no other peaks for any new phases or intermetallic compounds were formed due to the rapid consolidation process (45 min) during the SPS technique.
The crystallite size was assessed by the classical Williamson–Hall method (FWHM) from the broadening of XRD peaks and using the following formula [37,38]:
β cos θ λ = k d + 2 ε ( 2 sin θ λ )
where β is the full width at half maximum height (FWHM), θ is the Bragg’s angle of the peak, λ is the wavelength of X-ray (0.15406 nm), K is a dimensionless shape factor (0.9), which depends on the material, d is the crystallite size, and ε is the microstrain.
Figure 3 shows the effect of ceramic ratio on the crystallite size. A clear reduction of the crystallite size with increasing wt.% of ceramic additives was observed. Al2O3 and TiC are ceramic materials that act as internal balls that reduce the particle size [39,40]. In addition, the SPS technique achieved the consolidation process, which is a rapid method for the sintering in which no chance for the grain growth of the particles occurs [41,42]. The crystallite size of pure copper was ~105 nm, whereas the crystallite size for the produced composites was in the range of 5–25 nm.

3.2. Densification

The density of composite material is the most important parameter, which significantly affects both physical and mechanical properties. The relative density is calculated and plotted in Figure 4. Relative density is the ratio of the measured and theoretical density of the sample. Measured density was determined by the Archimedes method, and the theoretical density was calculated from the simple rule of mixtures. Each percent of pure Cu and 3, 6, 9 wt.% TiC/Al2O3 (individual and hybrid) nanocomposite were tested by three samples, and the average results were obtained. It was observed that the relative density decreased with increasing reinforcement content for all composites, as shown in Figure 4. The maximum relative density (~96%) was achieved by adding 3 wt.% TiC to copper, whereas the minimum relative density (89%) was obtained for 9 wt.% Al2O3/copper composite. TiC (4.91 g/cm3) and Al2O3 (3.987 g/cm3) also have lower densities than Cu [7,35]. So, the addition of a light material to a denser one decreased the overall density of the prepared composites. This may be attributed to the presence of hard ceramic material with a high melting point into a ductile metal such as Cu that may hinder the high densification and increase the porosity content accompanied by the high fraction of reinforcement [7,43].
Ayman Elsayed et al. [11] studied experimental investigations for the synthesis of W–Cu nanocomposite through spark plasma sintering, and they concluded that using the SPS technique led to reaching a maximum of 90% relative density. On the other hand, a relative density of 98.1% was reached for Cu-Fe-Al2O3-MoS2 composite sintered using the SPS route [33]. Moreover, Babapoor et al. [44] investigated the effects of spark plasma sintering temperature on the densification of TiC. They reached a relative density of 99.4% at 1900 °C for 7 min under 40 MPa using the TiC powder with a mean particle size of 7 μm. They also suggested that there is an optimum temperature for reaching the maximum density.

3.3. Microstructure Analysis

Figure 5 shows the FE-SEM micrograph of TiC-reinforced copper composite using 3%, 6%, and 9% TiC addition to Cu. Two phases are observed; the dark-gray phase represents the Cu matrix, and the black phase is the TiC particles. For 3 wt.% samples, TiC and Al2O3 particles are concentrated along the grain boundaries in a chain form, while 6 and 9 wt.% samples were homogeneously distributed all over the Cu matrix. This may be attributed to the suitable mechanical milling parameters and good SPS technique applied. The SPS technique leads to a finer structure compared with traditional routes [18,44].
The SEM investigation of Cu/Al2O3 microstructure is shown in Figure 6. Two phases are observed; the dark-gray phase represents the Cu matrix while the white phase represents the Al2O3 particles. The dispersion of Al2O3 inside the copper matrix is observed for Cu/3% Al2O3 with a little agglomeration of Al2O3 reinforcement. On the other hand, white areas of agglomerated alumina reinforcement are revealed within the Cu/6% Al2O3 matrix grain boundaries, whereas very fine particles are dispersed within the grain interior. Moreover, the SEM of Cu/9% Al2O3 composite shows that most of the Al2O3 nanoparticles are agglomerated along grain boundaries, and a small percentage are dispersed with the grains. Some authors have also concluded that increasing agglomeration steadily occurs, along with increasing the weight percentage of reinforcement [9,15].
The combination of both TiC and Al2O3 for reinforcing copper (Cu/hybrid nanocomposite) with point analysis EDS is shown in Figure 7a,b. The homogeneous distribution of TiC nanoparticles is predominant, while the agglomeration of some Al2O3 is observed along grain boundaries. More agglomeration of alumina particles along grain boundaries is observed with increasing hybrid percentage (TiC and Al2O3).

3.4. Mechanical Strength

Figure 8 represents the stress–strain curves for pure copper and TiC/Al2O3-reinforced copper matrix composites, while the key mechanical properties obtained from the compression test are plotted in Figure 9. The compressed samples are photographed in Figure 10. The addition of TiC enhanced the compression strength of copper and reached its maximum compression strength of ~852 N/mm2 at 9 wt.% TiC, whereas Al2O3 additions exhibited a dramatic effect on the Cu strength. Increasing Al2O3 from 3 to 6 wt.% increased the Cu strength, but a clear failure of strength is noticed at 9 wt.% ratios at which a minimum value of compression strength 367.8 N/mm2 resulted. After the compression test, the Cu-6% Al2O3 and 9% Al2O3 samples were destroyed (see Figure 10).
Moreover, the strength of hybrid composites increased firstly with increasing the percentage of reinforcement up to 6% and then slightly decreased. The extreme drop in the compression strength of the Cu/9% Al2O3 composite may be attributed to particle-to-particle contact resulting from ceramic particle agglomeration (see Figure 6). The high compression strength of TiC-strengthened Cu prepared by the SPS route compared with Cu/Al2O3 composite with the same wt.% is attributed to a combined effect of ultrafine grain (UFG) structure by the Hall–Petch mechanism and the obstruction of dislocation movement by nanoscale ceramic particles in the grain interior by the Orowan mechanism [8,43].
The effect of ceramic additions on the hardness of copper is illustrated in Figure 11. The hardness steadily increases with increasing the wt.% of reinforcement for synthesized composites. Cu/9% Al2O3 obtained the maximum hardness (211 HV), whereas two composites that obtained the minimum hardness value of 112 HV are Cu/3% TiC and Cu/3% hybrid. Many reasons could explain this. The first is that adding high-hardness and high-strength ceramic materials such as TiC and Al2O3 on the ductile Cu matrix increases the overall hardness. The second is that the addition of nanomaterials with the incorporation of nanoparticles between the Cu particles improves the hardness as a grain reinforcement takes place accordingly.

3.5. Strengthening Criteria

Increasing the strength of metallic materials is based on two competing factors. The first is work hardening, and the second is dynamic softening. Work hardening is caused by dislocations, multiplication, pileup, and tangle. Dynamic softening is caused by dislocations, rearrangement, and interactions. In the present work, the compression test and hardness measurement are carried out at room temperature, which is why the dynamic softening factor would not be probable, and the work hardening mechanism would affect the enhancement of strength and hardness. This is true for pure metal and alloys, unlike the composite materials where the contribution of ceramic additions to the matrix to enhance the properties should be considered.
Moreover, it is worth mentioning that some authors have considered the strengthening mechanisms in ceramic-reinforced composites [6,7,15,40,42,43,44]. The addition of TiC nanoparticles to the Cu matrix retained grain growth during sintering due to the peening effect of TiC for grain boundary movement and the strengthening effect of dispersed TiC in the Cu matrix grains where a mismatch of coefficient of thermal expansion is present [45] (see Figure 12). Furthermore, increasing the TiC fraction increased the strength and hardness of Cu-TiC composites in the present work. Another strengthening mechanism of TiC dispersion is the Orowan mechanism, especially at low fractions of TiC [6]. Compared to the other two cases, the composites with 3% and 6% Al2O3 gave the highest yield strength. This is a signal of increasing material strength with decreasing ductility. This may be attributed to the good adhesion between Cu matrix and TiC nanoparticles than between Cu and Al2O3 [46]. An extreme drop in the Cu-Al2O3 composite strength is noticed at 9 wt.% of Al2O3. This unexpected behavior may be attributed to the agglomeration of some alumina particles in the Cu matrix, increasing the chance for particle-to-particle contact (see Figure 13). A balanced behavior was observed with the hybrid (TiC/Al2O3) additions to the Cu matrix in which the combined effects of both ceramics are clear.
Reinforcing the Cu matrix, which is ductile in nature with two types of ceramic materials—ceramic carbide (TiC) and ceramic oxide (Al2O3)—helps to improve the mechanical properties of the Cu matrix. Both TiC and Al2O3 are at the nanoscale; therefore, by high ball milling, they filled the interstitial voids between Cu particles. As a consequence, the strengthening effect of both of them is distributed all over the Cu matrix. The hardness estimation test increased as the nano-ceramic hard particles were increased. This can also be explained by the resistance of the hard ceramic particles to the indenter from greater depth in the Cu-composite surface.
Consequently, the hardness is enhanced [46,47]. For the compression test, the presence of the nano-ceramic particles dispersed formally in the Cu matrix prevents the dislocation of the particles. In addition, as these hybrid reinforcements are at the nanoscale, they fill the voids; consequently, the strength of samples is increased [47,48,49].
Table 2 shows a comparison between the present study and previous work used to fabricate copper composites reinforced with alumina and titanium carbide nanoparticles. In this work, different concentrations of nano alumina and/or nano titanium carbide particles were used as a reinforcement material to the Cu matrix manufactured by the SPS technique. This work is compared with the same composites prepared by traditional sintering, vacuum sintering, hot pressing, and hot extrusion. The table shows that the composites produced by the SPS technique have the best mechanical properties compared with the other consolidation techniques.

4. Conclusions

Cu-(TiC and/or Al2O3) nanocomposites were synthesized successfully using mechanical milling followed by the spark plasma sintering (SPS) technique.
The density decreased with increasing percentages of the nano reinforcements (TiC and/or Al2O3). A maximum relative density of ~96% was achieved with the addition of 3 wt.% TiC to copper, whereas the minimum relative density (89%) was obtained by adding 9 wt.% Al2O3 to copper.
Agglomerated areas of Al2O3 nanoparticles around grain boundaries were observed, and increased with increasing Al2O3 fractions that, in turn, adversely affect the mechanical properties.
The compression strength of Cu/TiC increased with increasing the TiC fraction, and a maximum value of 851.99 N/mm2 was obtained by Cu/9% TiC. A dramatic behavior was observed for Cu/Al2O3 composites that gave the minimum compressive strength of 367.8 N/mm2 resulted at 9% Al2O3.
The maximum hardness of 211 HV was obtained by Cu/9% Al2O3, whereas two types of composites obtained the minimum hardness value of 112 HV: the Cu/3% TiC and Cu/3% hybrid.

Author Contributions

F.S.H.: investigation, writing—original draft and preparation; O.A.E.: Investigation, writing—original draft, review, and editing; A.R.S.E.: conceptualization and formal analysis; A.E.-N.: investigation and review; A.E.: methodology, review, and editing; A.K.E.: review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are included in the article. Any further requested information can be addressed to the corresponding author.

Acknowledgments

The authors acknowledge the kind support by Katsuyoshi Kondoh, Composite Materials Processing Lab., Osaka University, for providing the spark plasma sintering machine in his laboratory to carry out the consolidation of the composites and thank the researchers and technicians of the Central Metallurgical R & D Institute (CMRDI) in Cairo, Egypt for their collaboration.

Conflicts of Interest

The authors declare no potential conflict of interest concerning this article’s research, authorship, and/or publication.

References

  1. Daghigh, R.; Oramipoor, H.; Shahidian, R. Improving the Performance and Economic Analysis of Photovoltaic Panel Using Copper Tubular-Rectangular Ducted Heat Exchanger. Renew. Energy 2020, 156, 1076–1088. [Google Scholar] [CrossRef]
  2. Abyzov, A.M.; Kidalov, S.V.; Shakhov, F.M. High Thermal Conductivity Composite of Diamond Particles with Tungsten Coating in a Copper Matrix for Heat Sink Application. Appl. Therm. Eng. 2012, 48, 72–80. [Google Scholar] [CrossRef]
  3. Xiao, Y.; Yao, P.; Zhou, H.; Zhang, Z.; Gong, T.; Zhao, L.; Deng, M. Investigation on Speed-Load Sensitivity to Tribological Properties of Copper Metal Matrix Composites for Braking Application. Metals 2020, 10, 889. [Google Scholar] [CrossRef]
  4. Cao, G.; Konishi, H.; Li, X. Mechanical Properties and Microstructure of Mg/SiC Nanocomposites Fabricated by Ultrasonic Cavitation Based Nanomanufacturing. J. Manuf. Sci. Eng. Trans. ASME 2008, 130, 0311051–0311056. [Google Scholar] [CrossRef]
  5. Kaftelen, H.; Ünlü, N.; Göller, G.; Lütfi Öveolu, M.; Henein, H. Comparative Processing-Structure-Property Studies of Al-Cu Matrix Composites Reinforced with TiC Particulates. Compos. Part A Appl. Sci. Manuf. 2011, 42, 812–824. [Google Scholar] [CrossRef]
  6. Morris, D.G. The Origins of Strengthening in Nanostructured Metals and Alloys. Rev. Metal. 2010, 46, 173–186. [Google Scholar] [CrossRef] [Green Version]
  7. Sadoun, A.M.; Mohammed, M.M.; Fathy, A.; El-Kady, O.A. Effect of Al2O3 Addition on Hardness and Wear Behavior of Cu-Al2O3 Electro-Less Coated Ag Nanocomposite. J. Mater. Res. Technol. 2020, 9, 5024–5033. [Google Scholar] [CrossRef]
  8. Zhang, Z.; Chen, D.L. Contribution of Orowan Strengthening Effect in Particulate-Reinforced Metal Matrix Nanocomposites. Mater. Sci. Eng. A 2008, 483–484, 148–152. [Google Scholar] [CrossRef]
  9. Chandrakanth, R.G.; Rajkumar, K.; Aravindan, S. Fabrication of Copper-TiC-Graphite Hybrid Metal Matrix Composites through Microwave Processing. Int. J. Adv. Manuf. Technol. 2010, 48, 645–653. [Google Scholar] [CrossRef]
  10. Fathy, A.; El-Kady, O. Thermal Expansion and Thermal Conductivity Characteristics of Cu-Al2O3 Nanocomposites. Mater. Des. 2013, 46, 355–359. [Google Scholar] [CrossRef]
  11. Elsayed, A.; Li, W.; El Kady, O.A.; Daoush, W.M.; Olevsky, E.A.; German, R.M. Experimental Investigations on the Synthesis of W-Cu Nanocomposite through Spark Plasma Sintering. J. Alloys Compd. 2015, 639, 373–380. [Google Scholar] [CrossRef]
  12. Shyu, R.F.; Ho, C.T. In Situ Reacted Titanium Carbide-Reinforced Aluminum Alloys Composite. J. Mater. Process. Technol. 2006, 171, 411–416. [Google Scholar] [CrossRef]
  13. Ni, J.; Li, J.; Luo, W.; Han, Q.; Yin, Y.; Jia, Z.; Huang, B.; Hu, C.; Xu, Z. Microstructure and Properties of In-Situ TiC Reinforced Copper Nanocomposites Fabricated via Long-Term Ball Milling and Hot Pressing. J. Alloys Compd. 2018, 755, 24–28. [Google Scholar] [CrossRef]
  14. Wang, F.; Li, Y.; Wang, X.; Koizumi, Y.; Kenta, Y.; Chiba, A. In-Situ Fabrication and Characterization of Ultrafine Structured Cu-TiC Composites with High Strength and High Conductivity by Mechanical Milling. J. Alloys Compd. 2016, 657, 122–132. [Google Scholar] [CrossRef]
  15. Zhao, Q.; Gan, X.; Zhou, K. Enhanced Properties of Carbon Nanotube-Graphite Hybrid-Reinforced Cu Matrix Composites via Optimization of the Preparation Technology and Interface Structure. Powder Technol. 2019, 355, 408–416. [Google Scholar] [CrossRef]
  16. Akbarpour, M.R.; Mousa Mirabad, H.; Khalili Azar, M.; Kakaei, K.; Kim, H.S. Synergistic Role of Carbon Nanotube and SiCn Reinforcements on Mechanical Properties and Corrosion Behavior of Cu-Based Nanocomposite Developed by Flake Powder Metallurgy and Spark Plasma Sintering Process. Mater. Sci. Eng. A 2020, 786, 139395. [Google Scholar] [CrossRef]
  17. Akbarpour, M.R.; Alipour, S.; Farvizi, M.; Kim, H.S. Mechanical, Tribological and Electrical Properties of Cu-CNT Composites Fabricated by Flake Powder Metallurgy Method. Arch. Civ. Mech. Eng. 2019, 19, 694–706. [Google Scholar] [CrossRef]
  18. Kumar, R.; Chaubey, A.K.; Bathula, S.; Jha, B.B.; Dhar, A. Synthesis and Characterization of Al2O3-TiC Nano-Composite by Spark Plasma Sintering. Int. J. Refract. Met. Hard Mater. 2016, 54, 304–308. [Google Scholar] [CrossRef]
  19. Mohammadzadeh, A.; Akbarpour, M.R.; Heidarzadeh, A. Production of Nanostructured Copper Powder: Microstructural Assessments and Modeling. Mater. Res. Express 2018, 5, 065050. [Google Scholar] [CrossRef]
  20. Akbarpour, M.R.; Alipour, S. Wear and Friction Properties of Spark Plasma Sintered SiC/Cu Nanocomposites. Ceram. Int. 2017, 43, 13364–13370. [Google Scholar] [CrossRef]
  21. Hosseini, S.A.; Ranjbar, K.; Dehmolaei, R.; Amirani, A.R. Fabrication of Al5083 Surface Composites Reinforced by CNTs and Cerium Oxide Nano Particles via Friction Stir Processing. J. Alloys Compd. 2015, 622, 725–733. [Google Scholar] [CrossRef]
  22. Fono-Tamo, R.S.; Tien-Chien, J.; Akinlabi, E.T.; Sanusi, K.O. Surface Characteristics of Stainless Steel Powder in Magnesium Substrate: A Friction Stir Processed Composite. In Proceedings of the IEEE 10th International Conference on Mechanical and Intelligent Manufacturing Technologies (ICMIMT 2019), Cape Town, South Africa, 15–17 February 2019; pp. 10–14. [Google Scholar] [CrossRef]
  23. Mahmoud, E.R.I.; Ikeuchi, K.; Takahashi, M. Fabrication of SiC Particle Reinforced Composite on Aluminium Surface by Friction Stir Processing. Sci. Technol. Weld. Join. 2008, 13, 607–618. [Google Scholar] [CrossRef]
  24. Tinubu, O.O.; Das, S.; Dutt, A.; Mogonye, J.E.; Ageh, V.; Xu, R.; Forsdike, J.; Mishra, R.S.; Scharf, T.W. Friction Stir Processing of A-286 Stainless Steel: Microstructural Evolution during Wear. Wear 2016, 356–357, 94–100. [Google Scholar] [CrossRef] [Green Version]
  25. Escobar, J.D.; Velásquez, E.; Santos, T.F.A.; Ramirez, A.J.; López, D. Improvement of Cavitation Erosion Resistance of a Duplex Stainless Steel through Friction Stir Processing (FSP). Wear 2013, 297, 998–1005. [Google Scholar] [CrossRef]
  26. Rathee, S.; Maheshwari, S.; Siddiquee, A.N.; Srivastava, M. Distribution of Reinforcement Particles in Surface Composite Fabrication via Friction Stir Processing: Suitable Strategy. Mater. Manuf. Process. 2018, 33, 262–269. [Google Scholar] [CrossRef]
  27. Wen., H. Processing, Microstructure, Mechanical Behavior and Deformation Mechanisms of Bulk Nanostructured Copper and Copper Alloys; University of California: Davis, CA, USA, 2012. [Google Scholar]
  28. Deng, H.; Yi, J.; Xia, C.; Yi, Y. Improving the Mechanical Properties of Carbon Nanotube-Reinforced Pure Copper Matrix Composites by Spark Plasma Sintering and Hot Rolling. Mater. Lett. 2018, 210, 177–181. [Google Scholar] [CrossRef]
  29. Kim, K.T.; Cha, S., II; Hong, S.H.; Hong, S.H. Microstructures and Tensile Behavior of Carbon Nanotube Reinforced Cu Matrix Nanocomposites. Mater. Sci. Eng. A 2006, 430, 27–33. [Google Scholar] [CrossRef]
  30. Asgharzadeh, H.; Eslami, S. Effect of Reduced Graphene Oxide Nanoplatelets Content on the Mechanical and Electrical Properties of Copper Matrix Composite. J. Alloys Compd. 2019, 806, 553–565. [Google Scholar] [CrossRef]
  31. Deng, H.; Yi, J.; Xia, C.; Yi, Y. Mechanical Properties and Microstructure Characterization of Well-Dispersed Carbon Nanotubes Reinforced Copper Matrix Composites. J. Alloys Compd. 2017, 727, 260–268. [Google Scholar] [CrossRef]
  32. German, R.M. Particulate Composites; Springer International Publishing: Cham, Switzerland, 2016; ISBN 978-3-319-29915-0. [Google Scholar]
  33. Nautiyal, H.; Srivastava, P.; Khatri, O.P.; Mohan, S.; Tyagi, R. Wear and Friction Behavior of Copper Based Nano Hybrid Composites Fabricated by Spark Plasma Sintering. Mater. Res. Express 2019, 6, 0850h2. [Google Scholar] [CrossRef]
  34. Oanh, N.T.H.; Viet, N.H.; Kim, J.C.; Kim, J.S. Synthesis and Characterization of Cu–TiC Nanocomposites by Ball Milling and Spark Plasma Sintering. Mater. Sci. Forum 2014, 804, 173–176. [Google Scholar] [CrossRef]
  35. Oanh, N.T.H.; Viet, N.H.; Kim, J.S.; Dudina, D.V. Structural Investigations of TiC–Cu Nanocomposites Prepared by Ball Milling and Spark Plasma Sintering. Metals 2017, 7, 123. [Google Scholar]
  36. Farías, I.; Olmos, L.; Jiménez, O.; Flores, M.; Braem, A.; Vleugels, J. Wear Modes in Open Porosity Titanium Matrix Composites with TiC Addition Processed by Spark Plasma Sintering. Trans. Nonferrous Met. Soc. China 2019, 29, 1653–1664. [Google Scholar] [CrossRef]
  37. Abdel-Aziem, W.; Hamada, A.; Makino, T.; Hassan, M. Microstructural Evolution during Extrusion of Equal Channel Angular-Pressed AA1070 Alloy in Micro/Mesoscale. Mater. Sci. Technol. 2020, 36, 1169–1177. [Google Scholar] [CrossRef]
  38. Zak, A.K.; Majid, W.A.; Abrishami, M.E.; Yousefi, R. X-Ray Analysis of ZnO Nanoparticles by Williamson–Hall and Size–Strain Plot Methods. Solid State Sci. 2011, 13, 251–256. [Google Scholar] [CrossRef]
  39. Fathy, A.; Elkady, O.; Abu-Oqail, A. Production and Properties of Cu-ZrO2 Nanocomposites. J. Compos. Mater. 2018, 52, 1519–1529. [Google Scholar] [CrossRef]
  40. Sohag, M.A.Z.; Gupta, P.; Kondal, N.; Kumar, D.; Singh, N.; Jamwal, A. Effect of Ceramic Reinforcement on the Microstructural, Mechanical and Tribological Behavior of Al-Cu Alloy Metal Matrix Composite. Mater. Today Proc. 2020, 21, 1407–1411. [Google Scholar] [CrossRef]
  41. Cavaliere, P. Spark Plasma Sintering of Materials: Advances in Processing and Applications. Spark Plasma Sinter. Mater. Adv. Process. Appl. 2019, 1–781. [Google Scholar]
  42. Pan, Y.; Xiao, S.Q.; Lu, X.; Zhou, C.; Li, Y.; Liu, Z.W.; Liu, B.W.; Xu, W.; Jia, C.C.; Qu, X.H. Fabrication, Mechanical Properties and Electrical Conductivity of Al2O3 Reinforced Cu/CNTs Composites. J. Alloys Compd. 2019, 782, 1015–1023. [Google Scholar] [CrossRef]
  43. Tu, J.P.; Wang, N.Y.; Yang, Y.Z.; Qi, W.X.; Liu, F.; Zhang, X.B.; Lu, H.M.; Liu, M.S. Preparation and Properties of TiB2 Nanoparticle Reinforced Copper Matrix Composites by in Situ Processing. Mater. Lett. 2002, 52, 448–452. [Google Scholar] [CrossRef]
  44. Babapoor, A.; Asl, M.S.; Ahmadi, Z.; Namini, A.S. Effects of Spark Plasma Sintering Temperature on Densification, Hardness and Thermal Conductivity of Titanium Carbide. Ceram. Int. 2018, 44, 14541–14546. [Google Scholar] [CrossRef]
  45. Yin, Z.; Huang, C.; Zou, B.; Liu, H.; Zhu, H.; Wang, J. Study of the Mechanical Properties, Strengthening and Toughening Mechanisms of Al2O3/TiC Micro-Nano-Composite Ceramic Tool Material. Mater. Sci. Eng. A 2013, 577, 9–15. [Google Scholar] [CrossRef]
  46. Peng, T.; Yan, Q.; Zhang, X.; Zhuang, Y. Role of Titanium Carbide and Alumina on the Friction Increment for Cu-Based Metallic Brake Pads under Different Initial Braking Speeds. Friction 2021, 9, 1543–1557. [Google Scholar] [CrossRef]
  47. Venkateswarlu, M.; Kumar, M.A.; Reddy, K.H.C. Thermal Behavior of Spark Plasma Sintered Ceramic Matrix-Based Nanocomposites. J. Bio-Tribo-Corrosion 2020, 6, 6013–6028. [Google Scholar] [CrossRef]
  48. Wagih, A.; Abu-Oqail, A.; Fathy, A. Effect of GNPs Content on Thermal and Mechanical Properties of a Novel Hybrid Cu-Al2O3/GNPs Coated Ag Nanocomposite. Ceram. Int. 2019, 45, 1115–1124. [Google Scholar] [CrossRef]
  49. Kumar, E.G.; Ahasan, M.; Venkatesh, K.; Sastry, K.S.B.S.V.S. Design, Fabrication of Powder Compaction Die and Sintered Behavior of Copper Matrix Hybrid Composite. Int. Res. J. Eng. Technol. 2018, 5, 876–882. [Google Scholar]
  50. Akkaş, M.; Islak, S.; Özorak, C. Corrosion and Wear Properties of Cu-TiC Composites Produced by Hot Pressing Technique. Celal Bayar Üniversitesi Fen Bilim. Derg. 2018, 14, 465–469. [Google Scholar] [CrossRef]
  51. Palma, R.H.; Sepúlveda, A.H.; Espinoza, R.A.; Montiglio, R.C. Performance of Cu-TiC Alloy Electrodes Developed by Reaction Milling for Electrical-Resistance Welding. J. Mater. Process. Technol. 2005, 169, 62–66. [Google Scholar] [CrossRef]
  52. Akhtar, F.; Askari, S.J.; Shah, K.A.; Du, X.; Guo, S. Microstructure, Mechanical Properties, Electrical Conductivity and Wear Behavior of High Volume TiC Reinforced Cu-Matrix Composites. Mater. Charact. 2009, 60, 327–336. [Google Scholar] [CrossRef]
  53. Wang, F.; Li, Y.; Yamanaka, K.; Wakon, K.; Harata, K.; Chiba, A. Influence of Two-Step Ball-Milling Condition on Electrical and Mechanical Properties of TiC-Dispersion-Strengthened Cu Alloys. Mater. Des. 2014, 64, 441–449. [Google Scholar] [CrossRef]
  54. Öksüz, K.E.; Şahin, Y. Microstructure and Hardness Characteristics of Al2O3-B4C Particle-Reinforced Cu Matrix Composites. Acta Phys. Pol. A 2016, 129, 650–652. [Google Scholar] [CrossRef]
  55. Panda, S.; Dash, K.; Ray, B.C. Processing and Properties of Cu Based Micro- and Nano-Composites. Bull. Mater. Sci. 2014, 37, 227–238. [Google Scholar] [CrossRef] [Green Version]
  56. Orolínová, M.; Ďurišin, J.; Ďurišinová, K.; Danková, Z.; Ďurišin, M. Effect of Microstructure on Properties of Cu-Al2O3 Nanocomposite. Chem. Mater. Eng. 2013, 1, 60–67. [Google Scholar] [CrossRef]
  57. Fathy, A.; Shehata, F.; Abdelhameed, M.; Elmahdy, M. Compressive and Wear Resistance of Nanometric Alumina Reinforced Copper Matrix Composites. Mater. Des. 2012, 36, 100–107. [Google Scholar] [CrossRef]
  58. Dash, K.; Ray, B.C.; Chaira, D. Synthesis and Characterization of Copper-Alumina Metal Matrix Composite by Conventional and Spark Plasma Sintering. J. Alloys Compd. 2012, 516, 78–84. [Google Scholar] [CrossRef]
  59. Ritasalo, R.; Liu, X.W.; Söderberg, O.; Keski-Honkola, A.; Pitkänen, V.; Hannula, S.P. The Microstructural Effects on the Mechanical and Thermal Properties of Pulsed Electric Current Sintered Cu-Al2O3 Composites. Procedia Eng. 2011, 10, 124–129. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of Cu with TiC and Al2O3 and (TiC + Al2O3) hybrid nano reinforcement composites.
Figure 1. Schematic diagram of Cu with TiC and Al2O3 and (TiC + Al2O3) hybrid nano reinforcement composites.
Crystals 11 01081 g001
Figure 2. XRD patterns of composites after (SPS) process.
Figure 2. XRD patterns of composites after (SPS) process.
Crystals 11 01081 g002
Figure 3. The effect of ceramic additions on the crystallite size.
Figure 3. The effect of ceramic additions on the crystallite size.
Crystals 11 01081 g003
Figure 4. The effect of reinforcement fraction on the density of the produced composites. (a) Cu/TiC composites, (b) Cu/Al₂O₃ composites, (c) Cu/hybrid composite, and (d) Comparison of relative density between three series.
Figure 4. The effect of reinforcement fraction on the density of the produced composites. (a) Cu/TiC composites, (b) Cu/Al₂O₃ composites, (c) Cu/hybrid composite, and (d) Comparison of relative density between three series.
Crystals 11 01081 g004
Figure 5. SEM of Cu/TiC nanocomposite with TiC percentage of 3, 6, and 9% prepared using spark plasma sintered route.
Figure 5. SEM of Cu/TiC nanocomposite with TiC percentage of 3, 6, and 9% prepared using spark plasma sintered route.
Crystals 11 01081 g005
Figure 6. SEM of Cu/Al2O3 nanocomposite with Al2O3 percentage of 3, 6, and 9%.
Figure 6. SEM of Cu/Al2O3 nanocomposite with Al2O3 percentage of 3, 6, and 9%.
Crystals 11 01081 g006
Figure 7. (a) SEM of Cu/hybrid nanocomposite with (TiC+Al2O3) weight percentage of 3, 6, and 9% prepared using the spark plasma sintered route; (b) point analysis for Cu-reinforced hybrid ceramic matrix composite containing 9 wt.% hybrid ratio.
Figure 7. (a) SEM of Cu/hybrid nanocomposite with (TiC+Al2O3) weight percentage of 3, 6, and 9% prepared using the spark plasma sintered route; (b) point analysis for Cu-reinforced hybrid ceramic matrix composite containing 9 wt.% hybrid ratio.
Crystals 11 01081 g007
Figure 8. Compressive stress–strain curve for pure Cu and Cu-reinforced composites prepared by the spark plasma sintering route, (a) Cu/TiC composites, (b) Cu/Al2O3 composites, and (c) Cu/hybrid composite.
Figure 8. Compressive stress–strain curve for pure Cu and Cu-reinforced composites prepared by the spark plasma sintering route, (a) Cu/TiC composites, (b) Cu/Al2O3 composites, and (c) Cu/hybrid composite.
Crystals 11 01081 g008
Figure 9. Effect of nanosized TiC and Al2O3 contents on the mechanical properties of the Cu-based composites prepared by spark plasma sintering.
Figure 9. Effect of nanosized TiC and Al2O3 contents on the mechanical properties of the Cu-based composites prepared by spark plasma sintering.
Crystals 11 01081 g009
Figure 10. Compressed test samples for pure Cu and Cu matrix nanocomposites.
Figure 10. Compressed test samples for pure Cu and Cu matrix nanocomposites.
Crystals 11 01081 g010
Figure 11. A diagram of the measured hardness for copper-based nanoceramics.
Figure 11. A diagram of the measured hardness for copper-based nanoceramics.
Crystals 11 01081 g011
Figure 12. Spot white arrows indicate TiC around Cu grains.
Figure 12. Spot white arrows indicate TiC around Cu grains.
Crystals 11 01081 g012
Figure 13. SEM showing alumina distribution behavior in the Cu matrix.
Figure 13. SEM showing alumina distribution behavior in the Cu matrix.
Crystals 11 01081 g013
Table 1. Composition of the prepared specimens and their contents in the Cu matrix.
Table 1. Composition of the prepared specimens and their contents in the Cu matrix.
Materials after Sintering by (SPS)Composition
MatrixReinforcement
Cu [wt.%]TiC [wt.%]Al2O3 [wt.%]
0Pure copper100--------------
ICu-3 wt.% TiC973-------
Cu-6 wt.% TiC946-------
Cu-9 wt.% TiC919-------
IICu-3 wt.% Al2O397-------3
Cu-6 wt.% Al2O394-------6
Cu-9 wt.% Al2O391-------9
IIICu-1.5 wt.% TiC and 1.5 wt.% Al2O3971.51.5
Cu-3 wt.% TiC and 3 wt.% Al2O39433
Cu-4.5 wt.% TiC and 4.5 wt.% Al2O3914.54.5
Table 2. Comparing the present study with literature data of previous investigations.
Table 2. Comparing the present study with literature data of previous investigations.
CompositeMethodDensity, (g/cm3)Ultimate Stress,
(MPa)
Yield Stress,
(MPa)
Elongation, (%)Hardness,
(HV)
Ref. No
Pure copperSPS at 950 °C97N/A127.151.6981[Present study]
Cu-3 wt.% TiC96741.473133.97111.9
Cu-3 wt.% Al2O395587.435006.17149
Cu-1.5 wt.% TiC and 1.5 wt.% Al2O395414.86317.465.25112
Cu-5 wt.% TiCHot Press at 700 °C93.3N/AN/AN/A67.3[50]
Cu-5 vol.%TiCHot extrusionN/AN/AN/AN/A112[51]
Cu-5 vol.% TiCSPSN/A712661N/A221[14]
Cu-77 vol.% TiCSintering in Vacuum Furnace at 900 °C93.4N/AN/AN/A544[52]
Cu-5.3 vol.%TiCSPSN/A602572N/A194[53]
Cu-3 wt.% Al2O3 with Coating AgSintered at 950 °C95.9N/AN/AN/A85[7]
Cu-10 vol.% Al2O3Sintered at 880 °C83.49N/AN/AN/A71[54]
Cu-3 vol.% Al2O3Sintered at 850 °C91.5350N/A0.5177[55]
Cu-5 vol.% Al2O3Sintered at 850 °C88550N/A0.46100
Cu-5 vol.% Al2O3Sintering H2 at 850 °C 5304502.5155[56]
Cu-2.7 wt.% Al2O3Sintered at 950 °C92.53460350N/A54.83[57]
Cu-5 vol.% Al2O3Conventional Sintering N₂84.3N/AN/AN/A49[58]
Cu-5 vol.% Al2O3Conventional Sintering Ar84.3N/AN/AN/A48
Cu-5 vol.% Al2O3Conventional Sintering H₂94.4N/AN/AN/A79
Cu-5 vol.% Al2O3SPS at 700 °C92.2N/AN/AN/A125
Cu-2.75 wt.% Al2O3Pulsed Electric Current
Sintered (PECS)
99.6N/AN/AN/A94.83[59]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hamid, F.S.; A. Elkady, O.; Essa, A.R.S.; El-Nikhaily, A.; Elsayed, A.; Eessaa, A.K. Analysis of Microstructure and Mechanical Properties of Bi-Modal Nanoparticle-Reinforced Cu-Matrix. Crystals 2021, 11, 1081. https://doi.org/10.3390/cryst11091081

AMA Style

Hamid FS, A. Elkady O, Essa ARS, El-Nikhaily A, Elsayed A, Eessaa AK. Analysis of Microstructure and Mechanical Properties of Bi-Modal Nanoparticle-Reinforced Cu-Matrix. Crystals. 2021; 11(9):1081. https://doi.org/10.3390/cryst11091081

Chicago/Turabian Style

Hamid, Fadel S., Omayma A. Elkady, A. R. S. Essa, A. El-Nikhaily, Ayman Elsayed, and Ashraf K. Eessaa. 2021. "Analysis of Microstructure and Mechanical Properties of Bi-Modal Nanoparticle-Reinforced Cu-Matrix" Crystals 11, no. 9: 1081. https://doi.org/10.3390/cryst11091081

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