Next Article in Journal
Design and Preparation of CNTs/Mg Layered Composites
Previous Article in Journal
Effect of Sodium Hypochlorite Concentration on Electronic Apex Locator Reliability
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tribological Behavior of Aluminum Hybrid Nanocomposites Reinforced with Alumina and Graphene Oxide

by
Abdul Samad Mohammed
1,2,*,
Omar Saad Aljebreen
1,
Abbas Saeed Hakeem
3,
Tahar Laoui
4,
Faheemuddin Patel
1 and
Mirza Murtuza Ali Baig
1
1
Department of Mechanical Engineering, King Fahd University of Petroleum and Minerals, Dhahran 31261, Saudi Arabia
2
Interdisciplinary Research Center for Advanced Materials, King Fahd University of Petroleum and Minerals, Dhahran 31261, Saudi Arabia
3
Interdisciplinary Research Center for Hydrogen & Energy Storage (IRC-HES), King Fahd University of Petroleum & Minerals, Dhahran 31261, Saudi Arabia
4
Department of Mechanical and Nuclear Engineering, University of Sharjah, Sharjah P.O. Box 27272, United Arab Emirates
*
Author to whom correspondence should be addressed.
Materials 2022, 15(3), 865; https://doi.org/10.3390/ma15030865
Submission received: 30 December 2021 / Revised: 17 January 2022 / Accepted: 20 January 2022 / Published: 23 January 2022

Abstract

:
Due to rapid technological advancements, the demand for lightweight materials with improved tribo-mechanical properties is continuously growing. The development of composite materials is one of the routes taken by researchers to meet these target properties. Aluminum (Al) is one of the most suitable materials used for developing composites for a wide range of applications because of its light weight, high conductivity, and high specific strength. In this study, aluminum hybrid nanocomposites with alumina (10 Vol% Al2O3) and varying loadings of graphene oxide (0.25, 0.5 and 1 wt% GO) were fabricated using the spark plasma sintering technique. The tribological properties of the developed hybrid composites were evaluated by conducting ball-on-disk wear tests at a normal load of 3N, with a sliding speed of 0.1 m/s, and for a sliding distance of 100 m. A 440C hardened stainless steel ball with a diameter of 6.3 mm and a hardness of 62 RC was used as a counterface. Scanning electron microscopy (SEM), elemental X-ray dispersive analysis (EDS), and optical profilometry were used to ascertain the involved wear mechanisms. The results revealed that Al-10 Vol%Vol% Al2O3-0.25 wt% GO hybrid nanocomposite showed an increase of 48% in the hardness, a reduction of 55% in the specific wear rate, and a reduction of 5% in COF compared with pure aluminum.

1. Introduction

The demand for lightweight and high strength materials has been on the rise, primarily due to an expansion of the aerospace and automobile industries [1,2,3,4,5]. Metallic materials are used extensively in various engineering applications for their excellent mechanical and thermal properties. However, they suffer from two major drawbacks, such as high wear and low corrosion resistance. Hence, as the world is moving towards more industrialization, an emphasis on energy conservation has led to many researchers looking for avenues to further improve the properties of metals with a special focus on improving the tribological and anti-corrosive properties of metals. However, the inadequacy of metals and alloys in providing both high strength and stiffness required for demanding engineering applications, coupled with two major drawbacks such as high wear and low corrosion resistance, has led to the development of metal matrix composites (MMCs). Metal matrix composites can be designed to possess qualities such as a low coefficient of friction, low wear, and good anti-corrosive properties, making them suitable for use in demanding engineering applications [6].
Aluminum is one of the lightweight materials extensively studied over the years because of its excellent thermal conductivity and high strength-to-weight ratio [7]. However, the use of pure Al is generally limited to simple engineering applications, which necessitated the development of aluminum-based alloys with enhanced mechanical, corrosion, and tribological properties for a wider range of applications [8,9,10,11]. Further improvements to the properties of aluminum-based materials were achieved through the development of aluminum metal-based matrix composites (Al-MMCs) by adding different reinforcements to achieve tailored property combinations [12,13]. Ceramic reinforcements such as silicon carbide [14,15,16], titania [17], alumina [18,19,20,21,22], a combination of different ceramics [23], and carbon-based materials such as carbon fibers [24,25], graphite [26,27], carbon nanotubes (CNTs) [28,29], graphene [30,31], and graphene oxide (GO) [32,33] were used as reinforcements to fabricate Al-MMCs with improved properties.
Aluminum possesses high corrosion resistance under the majority of service conditions but has inferior resistance to wear. This has prompted researchers to develop Al-MMCs with improved tribological properties by incorporating different ceramic and carbon-based reinforcements by employing various fabrication methods. Irrespective of the fabrication process, tribological properties of Al-MMCs with single reinforcement showed improvement over pure aluminum and its alloys, and hybrid reinforcement demonstrated better tribological properties over single reinforcement [34]. Al-MMCs prepared by incorporating 5 wt% of B4C particles as reinforcement using the stir casting method showed improved tribological properties. Sliding wear rate and coefficient of friction were found to increase with the load but decrease with velocity and sliding distance [35]. Stir cast Al-MMCs prepared with 0.5, 1, 1.5, 2 and 2.5 wt.% nano silicon carbide powder as reinforcement showed a reduction in wear rate and friction coefficient with increasing SiC content. This was attributed to the increase in hardness and the lubricating nature of the SiC particles that became detached and polished the worn-out surface of the samples [36]. Wear behavior of Al2O3 nanoparticles reinforced Al-MMCs prepared using the ultrasonic-assisted semisolid stirring (UASS) method showed a progressive decrease in wear rate with an increase in reinforced Al2O3 content from 1% to 7%; however, wear rate increased with 10% Al2O3 content, which was attributed to an increase in agglomeration of Al2O3 nanoparticles [37]. A similar improvement in the wear resistance was observed in studies with Al2O3 and ZrO2 nanoparticles as reinforcements to Al-MMCs prepared using uniaxial compaction followed by sintering. The improvement in wear resistance correlated with the increase in hardness, owing to the hardness of the ceramic particle reinforcements [38,39]. The effect of hybrid ceramic reinforcements on the tribological behavior of Al-MMCs was also investigated. Hybrid SiC-TiC-Al-MMCs prepared using liquid stir casting process showed an improvement in wear resistance with an increase in the TiC reinforcement content from 1 to 2.5 wt%, whereas SiC content was fixed at 1 wt% [40]. Al-MMCs prepared using carbon-based materials as reinforcements were shown to affect tribological behavior. Al-MMCs prepared using the spark plasma sintering (SPS) process with 2 Vol%Vol% graphene nanoplatelets (GNPs) showed a lower coefficient of friction but increased wear rate under room temperature and at 200 °C wear testing. The decrease in the coefficient of friction was attributed to the slippage of interlayers of GNP. In contrast, the poor wear resistance was attributed to the agglomeration of GNPs, which resulted in poor densification and insufficient sintering [41]. However, it was found that the addition of graphite particles as a reinforcement to SiC-graphite-reinforced hybrid Al-MMCs led to the reduction in friction coefficient and frictional heat produced during sliding wear due to the formation of solid lubrication film [42]. The addition of ceramic reinforcements to Al-MMCs was shown to affect the corrosion behavior as well. Al-MMCs prepared with SiC reinforcements produced by uniaxial compaction and hot extrusion showed improved corrosion resistance upon increasing SiC content and was attributed to the chemical inertness of SiC particles to a corrosive solution. Moreover, SiC-Al-MMCs exhibited higher wear resistance and lower friction coefficients when wear-tested under corrosive media as compared with pure aluminum [43].
From the above literature, it is clear that the addition of ceramic reinforcement, as well as the addition of carbon reinforcement, tends to affect the tribological behavior of the developed Al-MMCs. Moreover, the type of processing method affects the developed properties. Few studies were reported on aluminum composites reinforced with either Al2O3 or graphene nanoplatelets individually. The present work reports on the synthesis, mechanical properties, and tribological behavior of aluminum hybrid nanocomposites reinforced with Al2O3 and GO, prepared using the spark plasma sintering process.

2. Materials and Methods

Aluminum (Al) powder from the Alpha chemical company, Dartmouth, Canada, with a purity of 99.5% and an average particle size of 30 µm was used as the matrix material. Aluminum oxide (Al2O3: Purity of 99.88%, particle size of 300 nm, approximate surface area of 85–115 m2/g) was procured from Union Carbide Corporation for BUEHLER Ltd. Lake Bluff, IL, USA and used as the first reinforcement. Graphene oxide (GO) from the AD Nano Company, Shimoga, India, with 99% purity and surface area of 250 m2/g was used as the second reinforcement.

2.1. Fabrication of the Hybrid Nanocomposites

In a previous study [44], the Al2O3 content in the Al matrix in terms of density and hardness and GO content in terms of mechanical and thermal properties was comprehensively optimized. It was observed that Al-10 Vol%Vol% Al2O3 resulted in the highest hardness of 55.8 HV compared with the other tested loadings of 20 and 30 Vol%Vol%, respectively. The improvement in the properties of Al-10 Vol%Vol% Al2O3 was attributed to the uniform dispersion of Al2O3 within the Al matrix, as observed from the scanning electron microscopy images. Moreover, among the hybrid nanocomposites, Al-10 Vol% Al2O3-0.25 wt% GO exhibited the highest hardness of 63 HV, the highest compressive strength of 180 MPa, and the lowest thermal expansion of 14.82 ppm °C−1 compared with the pristine and Al-10 Vol% Al2O3 samples. Since the mechanical and thermal properties of the Al-10 Vol% Al2O3-X wt% GO hybrid nanocomposites were comprehensively evaluated in [44], the aim of the present study was defined as the evaluation of tribological properties of Al-10 Vol% Al2O3-X wt% GO, wherein X represents the different GO loadings (0.25, 0.5 and 1 wt%) in the hybrid nanocomposite. The detailed procedure of preparing the hybrid nanocomposites, including the ball milling parameters and the spark plasma sintering parameters, can be found in the earlier research [44] and is also described below briefly.

2.1.1. Sonication and Ball Milling to Disperse and Mix the Powders

To obtain a uniform dispersion, 10 Vol% Al2O3 nanoparticles were sonicated in ethanol for 1 h using a probe sonicator (Sonics VCX 750, Newtown, CT, USA) with a cycle amplitude of 45% (ON: 20 s/Off 5 s) at room temperature. The same procedure was repeated for breaking the agglomeration in GO. The dispersed Al2O3 was then added to the weighed quantity of Al powder for ball milling (HD/HDDM/01, Union process, Inc., Akron, OH, USA) in zirconia vials. Ball milling was carried out for 24 h at a rotational speed of 200 rpm, using zirconia balls (diameter = 5 mm) with a ball to powder ratio of 10:1 for homogeneous mixing. The milling process was carried out in an argon atmosphere to avoid oxidation of the powders and in the presence of 50 mL of ethanol to avoid any cold welding during the process. The mixed powders were then dried in an oven at a temperature of 80 °C for 12 h to evaporate ethanol completely. To prepare the hybrid powders, GO was added in different percentages of 0.25, 0.5, and 1 wt%, respectively, to the mixture of Al-10 Vol% Al2O3 and mixed for 24 h with the same milling parameters.

2.1.2. Spark Plasma Sintering Process

Circular samples of 20 mm diameter and a thickness of 6 mm for each formulation were produced using spark plasma sintering (SPS) equipment (SPS FCTsystem GmBH, Frankenblick, Germany) under a high vacuum. Table 1 shows the SPS parameters used for fabricating the samples.

2.2. Densification and Hardness Measurements

The density of the fabricated samples was measured by immersing the sample in deionized water according to Archimedes principles and rule of mixtures by using (Kern ABT weighing scale, 320 g capacity, Nordrhein-Westfalen, Germany). The hardness was measured using a Zwick hardness tester by using an applied load of 500 gf. An average of 10 readings is reported for each sample.

2.3. Friction and Wear Tests

Ball on disk wear experiments were conducted using a tribometer (UMT-3, Bellerica, MA, USA) to determine the friction and wear characteristics of the composites. The counterface used was a 440C stainless steel ball with a diameter of 6.3 mm and a hardness of 62 RC. Wear tests were conducted at a normal load of 3 N, with a sliding peed of 0.1 m/s for 5000 cycles corresponding to a sliding distance of 100 m. The wear volume loss was estimated using a 3D optical profilometer (GTK-A, Bruker, Bellerica, MA, USA). A scanning electron microscope (SEM: Tescan VEGA3, Brno, Czech Republic) with an electron X-ray dispersive (EDS) attachment was used to ascertain the involved wear mechanisms and the role of the wear debris particles. An optical microscope was used to record the counterface ball images to evaluate its wear.

3. Results

3.1. Density and Hardness of the Nanocomposites

Table 2 shows the densification of all the formulations, measured based on the Archimedes method. It was observed that the density of the samples reduced marginally with the addition of the reinforcements, with pure aluminum showing the highest density of 99.8% compared with the other compositions.
Figure 1 shows the hardness variation with the addition of Al2O3 and GO to the Al matrix. It can be observed from Figure 1 that the hardness increased significantly with the addition of Al2O3 from 32.80 HV to 55.83 HV. This is attributed to the inherent hardness of the Al2O3 particles and their uniform dispersion within the Al matrix, as reported in the earlier study [44]. Moreover, with the addition of 0.25 wt% of GO to the Al-10 Vol% Al2O3 composite, the hardness increased to 63.57 HV, after which the hardness decreased with increasing GO content. The increase in hardness with the addition of 0.25 wt% GO was attributed to the uniform dispersion of both fillers within the aluminum matrix without the formation of agglomerates. Moreover, evidence from the SEM images of the microstructure analysis of Al-10 Vol% Al2O3-0.25 wt% GO showed a very fine grain size which contributed to the increase in the hardness. However, the reduction in hardness of the hybrid composites with an increase in the content of GO was attributed to the formation of agglomerates and lower densification of the samples (Table 2) [44]. The variation of hardness with the addition of only GO and Al2O3 can also be seen from Figure 1. Clearly, the addition of 0.25 wt% GO resulted in a marginal increase in hardness compared with the addition of Al2O3 nanoparticles. This indicates that Al2O3 reinforcement is the main contributor for the improvement in the hardness of the hybrid nanocomposite.

3.2. Tribological Characterization of the Nanocomposite/Hybrid Samples

3.2.1. Specific Wear Rate of the Developed Nanocomposite/Hybrid Samples

Figure 2 shows the variation in the specific wear rate of the various samples tested under an applied load of 3 N and a sliding speed of 0.1 m/s for a duration of 5000 cycles corresponding to a sliding distance of 100 m. The specific wear rate was calculated by the ratio of the total wear volume loss to the product of sliding distance and the applied load. The wear volume loss was estimated by the product of the wear area, measured by analyzing the 2D profiles of the wear tracks recorded by using a 3D optical profilometer and the circumference of the wear track. Figure 3 shows a few typical 2D profiles as measured by the 3D optical profilometer for each tested sample. According to Figure 2, the specific wear rate of the aluminum composites (Al-0.25 wt% GO and Al-10 Vol% Al2O3) decreased as compared to pure aluminum, noting that the reduction in the specific wear rate for Al-10 Vol% Al2O3 was greater than that of Al-0.25 wt% GO. This could be attributed to the higher hardness of the Al-10 Vol% Al2O3 composite compared with that of the Al-0.25 wt% GO composite, as reported in Figure 1.
For the hybrid nanocomposites, the specific wear rate decreased slightly further upon adding 0.25 wt% GO to Al-10 Vol% Al2O3, owing to the improved hardness of the hybrid nanocomposite (Figure 1), mainly due to the hard alumina nanoparticles, and to the lubricious nature of GO. Moreover, as reported earlier [44], for the hybrid nanocomposite Al-10 Vol% Al2O3-0.25 wt% GO, uniform dispersion of both reinforcements within the aluminum matrix was observed, leading to an effective load transfer, which reults in an improvement in the wear resistance. However, as the content of GO increased (0.5 and 1 wt%), the specific wear rate of the hybrid nanocomposites increased significantly. This could be attributed to the reduction in hardness of the samples (Figure 1), lower densification (Table 1), and as reported earlier [44], the formation of agglomerates of the reinforcements within the aluminum matrix.

3.2.2. Friction Coefficient of the Developed Nanocomposites/Hybrid Samples

Figure 4 and Figure 5 show typical frictional graphs for the wear tests and the average coefficient of friction (COF) for all tested samples under a normal load of 3 N, a sliding speed of 0.1 m/s, and a sliding distance of 100 m. According to Figure 5, the average COF increased slightly for the Al-10 Vol% Al2O3 nanocomposites compared with pure Al, mainly due to the presence of the hard alumina particles. Moreover, the COF of Al-0.25 wt% GO was less than that of pure Al and the Al-10 Vol% Al2O3 nanocomposite, due to the lubricious nature of GO. However, for the hybrid nanocomposite, Al-10 Vol% Al2O3-0.25 wt% GO, the COF increased slightly as compared with the Al-0.25 wt% GO nanocomposites, due to the addition of hard alumina particles. The COF for the Al-10 Vol% Al2O3-0.25 wt% GO hybrid nanocomposite was lower than pure Al and Al-10 Vol% Al2O3 nanocomposites, due to the presence of lubricious GO. In fact, a further increase in the GO content led to a further decrease in COF. The main reason for adding GO to the hybrid nanocomposite was to reduce the COF, as confirmed by these results.

3.2.3. SEM/EDS Analysis

Figure 6 shows the SEM/EDS images of the wear tracks, and Figure 7 shows the optical images of the counterface ball after sliding against the tested samples at a load of 3 N, sliding speed of 0.1 m/s, and for a sliding distance of 100 m.
Pure Al displayed a high wear rate, as shown in Figure 2, and was confirmed by SEM image of the wear track shown in Figure 6a. It shows a very rough surface characterized by deep grooves and debris particles resulting from the ploughing of the surface during the wear test involving a two-body and a three-body abrasive wear mechanism coupled with severe adhesive wear (gouging) mechanism (Figure 6a). The high COF of 0.9 shown by Al (Figure 5) was attributed to the high adhesive and the deformation components of the friction coefficient. The counterface ball sliding against the samples also showed a large scar mark, as shown in Figure 7a.
However, on the addition of 0.25 wt% GO alone to the Al matrix, a reduction of 22% and 9% in the specific wear rate and the COF was observed, respectively. Figure 6b shows a smoother wear track as compared with the pure Al, with a few instances of adhesive wear along the edges. An increase in hardness was observed upon adding 0.25 wt% GO to Al matrix (Figure 1), which could be the reason for its better tribological performance in terms of increased wear resistance. The inherent lubricating properties of GO must have contributed to the reduction in the COF as compared with Al. Moreover, the counterface ball sliding against the Al-0.25 wt% GO also showed a scar mark of similar size, as shown by the ball sliding against the Al sample.
The addition of 10 Vol% Al2O3 alone to Al matrix resulted in a significant increase in the hardness, as shown in Figure 1. The Al-10 Vol% Al2O3 composite showed a significant decrease of 46% in the specific wear rate, accompanied by a 6% increase in the COF as compared with Al (Figure 2 and Figure 5). A close examination of the wear track showed the wear mechanism to be mostly abrasive in nature, with ploughing marks visible in the direction of sliding. Small traces of iron were detected on the wear track, which could have been transferred from the counterface ball during the test (Figure 6c). The scar mark on the counterface ball was slightly smaller (Figure 7c) compared with the one observed in the earlier cases. Thus, it can be inferred that the addition of Al2O3 to Al matrix contributed mainly to the improvement of the hardness of Al, leading to a significant decrease in the specific wear rate.
Upon adding both reinforcements to Al matrix to form a hybrid nanocomposite, an interesting trend was observed. A simultaneous addition of 10 Vol% Al2O3 and 0.25 wt% GO to Al resulted in an increase of 48% in the hardness, reduction of 55% in the specific wear rate and a reduction of 5% in COF compared with Al. Moreover, this hybrid nanocomposite exhibited a better wear resistance than Al-0.25 wt% GO and Al-10 Vol% Al2O3 composites (Figure 2). The SEM image analysis of the wear track revealed a smooth surface with traces of wear predominantly by plastic deformation. Moreover, the scar mark on the counterface ball was also relatively small compared with the other samples (Figure 7d). This excellent performance of the Al-10 Vol% Al2O3-0.25 wt% GO could be attributed to the improvement in the hardness due to the addition of both Al2O3 and GO particles, with the major contribution coming from the hard Al2O3 particles. The GO contributed to the reduction in the COF of the hybrid nanocomposite by forming a tribo film over the wear track, as revealed by EDS analysis of the wear track (Figure 6d).
A further increase in GO content (0.5 and 1 wt%) in Al-10 Vol% Al2O3 matrix resulted in a reduction in hardness, wear resistance and COF. The decrease in the hardness and density of the hybrid samples with the increased amount of GO was attributed to the formation of GO agglomerates, as reported in [44]. The formation of GO agglomerates resulted in a two-phase morphology, leading to a decrease in hardness, which in turn reduced the wear resistance of these hybrid samples (Figure 2). The examination of the SEM images of the wear track, performed after the wear test, indicated the primary wear mechanism to be abrasive wear due to ploughing, as groove marks were visible in the direction of sliding (Figure 6e,f). The scar marks on the counterface balls were also more significant than the balls that slid against the other samples (Figure 7e,f).

4. Summary and Comparative Analysis

Table 3 summarizes the results obtained in the current study for all tested samples in terms of relative density, hardness, wear rate and coefficient of friction. Clearly, Al-10 Vol% Al2O3-0.25 wt% GO nanocomposite showed the highest reduction of 55.65% in specific wear rate compared with Al.
Furthermore, the tribological performance of the Al-10 Vol% Al2O3-0.25 wt% GO sample showed the lowest wear rate in the current study, as compared with the results reported in the literature. Table 4 compares the tribological performance of the aluminum composites reinforced with Al2O3 and graphene, as reported in the literature, with the hybrid nanocomposite developed in the current study. It is worth noting that the developed aluminum hybrid nanocomposite reinforced with both Al2O3 and GO showed excellent tribological performance as compared with the performance of the composites reported in the literature.

5. Conclusions

Aluminum (Al) hybrid nanocomposites were fabricated using the spark plasma sintering technique. Alumina (Al2O3) and graphene oxide (GO) were used to reinforce the Al matrix. The tribological performance of the developed nanocomposites was evaluated. The following conclusions can be drawn from the study:
  • Al-10 Vol% Al2O3-0.25 wt% GO hybrid nanocomposite showed the maximum increase in hardness of 48.4% compared with Al. This significant increase was attributed to the inherent hard nature of Al2O3 nanoparticles and the uniform dispersion of both the reinforcements within the Al matrix.
  • Al-10 Vol% Al2O3-0.25 wt% GO hybrid nanocomposite showed the lowest specific wear rate. It exhibited the highest reduction of about 55.6% in the specific wear rate as compared with Al. This was attributed to the enhancement in the mechanical properties due to the reinforcements.
  • The most predominant wear mechanism was found to be abrasive wear due to plastic deformation, with smooth surface characteristics of the wear track with significantly fewer debris particles. The scar mark on the counterface ball sliding against the Al-10 Vol% Al2O3-0.25 wt% GO hybrid nanocomposite was also the smallest.

Author Contributions

Conceptualization, methodology, funding acquisition, project administration, supervision, validation, writing—original draft preparation, writing—review and editing, A.S.M. Formal analysis, Investigation, data curation, O.S.A. Investigation, writing—review and editing, A.S.H. Methodology, writing—review and editing, T.L. Data curation, writing—review and editing, F.P. Data curation, validation, M.M.A.B. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledge the financial support provided by the Deanship of Scientific Research (DSR) at King Fahd University of Petroleum & Minerals (KFUPM) through Project No. IN151009.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Clyne, T.W.; Withers, P.J. An Introduction to Metal Matrix Composites; Cambridge University Press: Cambridge, UK, 1993. [Google Scholar] [CrossRef]
  2. Kainer, K.U. Metal Matrix Composites: Custom-Made Materials for Automotive and Aerospace Engineering; John Wiley & Sons: Chichester, UK, 2006; p. 314. [Google Scholar]
  3. Lloyd, D.J. Particle reinforced aluminum and magnesium matrix composites. Int. Mater. Rev. 1994, 39, 1–23. [Google Scholar] [CrossRef]
  4. Lee, J.M.; Lee, S.K.; Hong, S.J.; Kwon, Y.N. Microstructures and thermal properties of A356/SiCp composites fabricated by liquid pressing method. Mater. Des. 2012, 37, 313–316. [Google Scholar] [CrossRef]
  5. Alidokht, S.A.; Zadeh, A.A.; Soleymani, S.; Assadi, H. Microstructure and tribological performance of an aluminum alloy based hybrid composite produced by friction stir processing. Mater. Des. 2011, 32, 2727–2733. [Google Scholar] [CrossRef]
  6. Torralba, J.; da Costa, C.; Velasco, F.J. P/M aluminum matrix composites: An overview. J. Mater. Process. Technol. 2003, 133, 203–206. [Google Scholar] [CrossRef]
  7. Chakrapani, P.; Suryakumari, T.S.A. Mechanical properties of aluminium metal matrix composites-A review. Mater. Today Proc. 2021, 45, 5960–5964. [Google Scholar] [CrossRef]
  8. Dursun, T.; Soutis, C. Recent developments in advanced aircraft aluminium alloys. Mater. Des. 2013, 56, 862–871. [Google Scholar] [CrossRef]
  9. Sevik, H.; Kurnaz, S.C. Properties of alumina particulate reinforced aluminum alloy produced by pressure die casting. Mater. Des. 2005, 27, 676–683. [Google Scholar] [CrossRef]
  10. Freitas, B.J.M.; Junior, A.M.J.; Zepon, G.; Kiminami, C.S.; Botta, W.J.; Bolfarini, C. Outstanding Tensile Ductility in High Iron-Containing Al-Si-Cu Alloys. Met. Mater. Trans. A 2020, 51, 2703–2710. [Google Scholar] [CrossRef]
  11. Hutchinson, C.R.; Ringer, S.P. Precipitation processes in Al-Cu-Mg alloys microalloyed with Si. Met. Mater. Trans. A 2000, 31, 2721–2733. [Google Scholar] [CrossRef]
  12. Dev, S.P.; Charoo, M.S. Aluminum metal matrix composites a review of reinforcement; Mechanical and tribological behavior. Int. J. Eng. Technol. 2018, 7, 117. [Google Scholar]
  13. Tjong, S.C. Processing and Deformation Characteristics of Metals Reinforced with Ceramic Nanoparticles, 2nd ed.; Tjong, J., Ed.; Elsevier: Oxford, UK, 2014; pp. 269–304. [Google Scholar]
  14. Tham, L.; Gupta, M.; Cheng, L. Effect of limited matrix–reinforcement interfacial reaction on enhancing the mechanical properties of aluminium–silicon carbide composites. Acta Mater. 2001, 49, 3243–3253. [Google Scholar] [CrossRef]
  15. Rana, R.; Purohit, R.; Soni, V.; Das, S. Characterization of Mechanical Properties and Microstructure of Aluminium Alloy-SiC Composites. Mater. Today Proc. 2015, 2, 1149–1156. [Google Scholar] [CrossRef]
  16. Song, M. Effects of volume fraction of SiC particles on mechanical properties of SiC/Al composites. Trans. Nonferrous Met. Soc. China 2009, 19, 1400–1404. [Google Scholar] [CrossRef]
  17. Kumar, C.A.V.; Rajadurai, J.S. Influence of rutile (TiO2) content on wear and microhardness characteristics of aluminium-based hybrid composites synthesized by powder metallurgy. Trans. Nonferrous Met. Soc. China 2016, 26, 63–73. [Google Scholar] [CrossRef]
  18. Ezatpour, H.; Torabi-Parizi, M.; Sajjadi, S.A. Microstructure and mechanical properties of extruded Al/Al2O3 composites fabricated by stir-casting process. Trans. Nonferrous Met. Soc. China 2013, 23, 1262–1268. [Google Scholar] [CrossRef]
  19. Zahid, G.; Azhar, T.; Musaddiq, M.; Rizvi, S.; Ashraf, M.; Hussain, N.; Iqbal, M. In situ processing and aging behaviour of an aluminium/Al2O3 composite. Mater. Des. 2011, 32, 1630–1635. [Google Scholar] [CrossRef]
  20. Razavi-Tousi, S.; Yazdani-Rad, R.; Manafi, S. Effect of volume fraction and particle size of alumina reinforcement on compaction and densification behavior of Al–Al2O3 nanocomposites. Mater. Sci. Eng. A 2010, 528, 1105–1110. [Google Scholar] [CrossRef]
  21. Jiang, X.-S.; Wang, N.-J.; Zhu, D.-G. Friction and wear properties of in-situ synthesized Al2O3 reinforced aluminum composites. Trans. Nonferrous Met. Soc. China 2014, 24, 2352–2358. [Google Scholar] [CrossRef]
  22. Rahimian, M.; Parvin, N.; Ehsani, N. The effect of production parameters on microstructure and wear resistance of powder metallurgy Al–Al2O3 composite. Mater. Des. 2010, 32, 1031–1038. [Google Scholar] [CrossRef]
  23. Shuvho, M.; Chowdhury, M.; Kchaou, M.; Roy, B.; Rahman, A.; Islam, M. Surface characterization and mechanical behavior of aluminum based metal matrix composite reinforced with nano Al2O3, SiC, TiO2 particles. Chem. Data Collect. 2020, 28, 100442. [Google Scholar] [CrossRef]
  24. Kumar, N.; Chittappa, H.C.; Ezhil Vannan, S. Development of Aluminium-Nickel Coated Short Carbon Fiber Metal Matrix Composites. Mater. Today Proc. 2018, 5, 11336–11345. [Google Scholar] [CrossRef]
  25. Wang, S.; Zhang, Y.; Sun, P.; Cui, Y.; Wu, G. Microstructure and Flexural Properties of Z-Pinned Carbon Fiber-Reinforced Aluminum Matrix Composites. Materials 2019, 12, 174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Sharma, P.; Sharma, S.; Khanduja, D. Effect of graphite reinforcement on physical and mechanical properties of alu-minum metal matrix composites. Part. Sci. Technol. 2016, 34, 17–22. [Google Scholar] [CrossRef]
  27. Schrecengost, T.R.; Shaw, B.A.; Wendt, R.G.; Moshier, W.C. Nonequilibrium Alloying of Graphite-Reinforced Aluminum Metal Matrix Composites. Corrosion 1993, 49, 842–849. [Google Scholar] [CrossRef]
  28. Jagannatham, M.; Chandran, P.; Sankaran, S.; Haridoss, P.; Nayan, N.; Bakshi, S.R. Tensile properties of carbon nanotubes reinforced aluminum matrix composites: A review. Carbon 2020, 160, 14–44. [Google Scholar] [CrossRef]
  29. Sridhar, I.; Narayanan, K.R. Processing and characterization of MWCNT reinforced aluminum matrix composites. J. Mater. Sci. 2009, 44, 1750–1756. [Google Scholar] [CrossRef]
  30. Munir, K.; Wen, C.; Li, Y. Graphene nanoplatelets-reinforced magnesium metal matrix nanocomposites with superior mechanical and corrosion performance for biomedical applications. J. Magnes. Alloy. 2019, 8, 269–290. [Google Scholar] [CrossRef]
  31. Pourmand, N.S.; Asgharzadeh, H. Aluminum Matrix Composites Reinforced with Graphene: A Review on Production, Microstructure, and Properties. Crit. Rev. Solid State Mater. Sci. 2019, 45, 289–337. [Google Scholar] [CrossRef]
  32. Lian, W.; Mai, Y.; Wang, J.; Zhang, L.; Liu, C.; Jie, X. Fabrication of graphene oxide-Ti3AlC2 synergistically reinforced copper matrix composites with enhanced tribological performance. Ceram. Int. 2019, 45, 18592–18598. [Google Scholar] [CrossRef]
  33. Dasari, B.; Morshed, M.; Nouri, J.M.; Brabazon, D.; Naher, S. Mechanical properties of graphene oxide reinforced aluminium matrix composites. Compos. Part B Eng. 2018, 145, 136–144. [Google Scholar] [CrossRef]
  34. Şahin, Y. Tribological behavior and wear surface analysis of metal-matrix composites. J. Mater. Sci. 1999, 34, 875–880. [Google Scholar] [CrossRef]
  35. Prabhakar, N.S.; Radhika, N.; Raghu, R. Analysis of Tribological Behavior of Aluminium/B4C Composite Under Dry Sliding Motion. Procedia Eng. 2014, 97, 994–1003. [Google Scholar] [CrossRef] [Green Version]
  36. Faisal, N.; Kumar, K. Mechanical and tribological behaviour of nano scaled silicon carbide reinforced aluminium composites. J. Exp. Nanosci. 2018, 13, S1–S13. [Google Scholar] [CrossRef]
  37. Jiang, J.; Xiao, G.; Che, C.; Wang, Y. Microstructure, Mechanical Properties and Wear Behavior of the Rheoformed 2024 Aluminum Matrix Composite Component Reinforced by Al2O3 Nanoparticles. Metals 2018, 8, 460. [Google Scholar] [CrossRef] [Green Version]
  38. Al-Qutub, A.; Allam, I.; Qureshi, T. Effect of sub-micron Al2O3 concentration on dry wear properties of 6061 aluminum based composite. J. Mater. Process. Technol. 2005, 172, 327–331. [Google Scholar] [CrossRef]
  39. Ramachandra, M.; Abhishek, A.; Siddeshwar, P.; Bharathi, V. Hardness and Wear Resistance of ZrO2 Nano Particle Reinforced Al Nanocomposites Produced by Powder Metallurgy. Procedia Mater. Sci. 2015, 10, 212–219. [Google Scholar] [CrossRef] [Green Version]
  40. Kumar, A.; Arafath, Y.; Gupta, P.; Kumar, D.; Hussain, C.M.; Jamwal, A. Microstructural and mechano-tribological behavior of Al reinforced SiC-TiC hybrid metal matrix composite. Mater. Today Proc. 2020, 21, 1417–1420. [Google Scholar] [CrossRef]
  41. Rengifo, S.; Zhang, C.; Harimkar, S.; Boesl, B.; Agarwal, A. Tribological Behavior of Spark Plasma Sintered Aluminum-Graphene Composites at Room and Elevated Temperatures. Technologies 2017, 5, 4. [Google Scholar] [CrossRef]
  42. Singh, J. Fabrication characteristics and tribological behavior of Al/SiC/Gr hybrid aluminum matrix composites: A review. Friction 2016, 4, 191–207. [Google Scholar] [CrossRef] [Green Version]
  43. Mosleh-Shirazi, S.; Akhlaghi, F.; Li, D.-Y. Effect of SiC content on dry sliding wear, corrosion and corrosive wear of Al/SiC nanocomposites. Trans. Nonferrous Met. Soc. China 2016, 26, 1801–1808. [Google Scholar] [CrossRef]
  44. Mohammed, A.; Alahmari, T.; Laoui, T.; Hakeem, A.; Patel, F. Mechanical and Thermal Evaluation of Aluminum Hybrid Nanocomposite Reinforced with Alumina and Graphene Oxide. Nanomaterials 2021, 11, 1225. [Google Scholar] [CrossRef] [PubMed]
  45. Ajit, S.D.; Jayakumar, J. Evaluation of mechanical and wear properties of aluminium-Al2O3 composite material. Int. Res. J. Eng. Technol. (IRJET) 2012, 5, 5–18. [Google Scholar]
  46. Bharath, V.; Madev, N.; Auradi, V.; Kori, S.A. Preparation of 6061Al-Al2O3 MMC’s by Stir Casting and Evaluation of Mechanical and Wear Properties. Procedia Mater. Sci. 2014, 6, 1658–1667. [Google Scholar] [CrossRef] [Green Version]
  47. Al-Salihi, H.A.; Judran, H.K. Effect of Al2O3 reinforcement nanoparticles on the tribological behaviour and mechanical properties of Al6061 alloy. AIMS Mater. Sci. 2020, 7, 486–498. [Google Scholar] [CrossRef]
  48. Tabandeh, K.; Meysam, O.; Emad, M.; Pradeep, R. and Pradeep, K.H. Tribological performance of self-lubricating aluminum matrix nanocomposites: Role of graphene nanoplatelets. Eng. Sci. Technol. Int. J. 2016, 19, 463–469. [Google Scholar]
  49. Xia, H.-M.; Zhang, L.; Zhu, Y.-C.; Li, N.; Sun, Y.-Q.; Zhang, J.-D.; Ma, H.-Z. Mechanical properties of graphene nanoplatelets reinforced 7075 aluminum alloy composite fabricated by spark plasma sintering. Int. J. Miner. Met. Mater. 2020, 27, 1295–1300. [Google Scholar] [CrossRef]
  50. Sekar, K.; Allesu, K.; Joseph, M.A. Mechanical and Wear Properties of Al–Al2O3 Metal Matrix Composites Fabricated by the Combined Effect of Stir and Squeeze Casting Method. Trans. Indian Inst. Met. 2015, 68, 115–121. [Google Scholar] [CrossRef]
  51. Aswin, S.; Soney, V. Tribology study of graphene-aluminium nanocomposites. Int. J. Mech. Prod. Eng. (IJMPE) 2014, 2, 29–33. [Google Scholar]
  52. Prashantha, K.H.G.; Anthony, X.M. Tribological Aspects of Graphene-Aluminum Nanocomposites; IntechOpen: London, UK, 2017; Available online: https://www.intechopen.com/chapters/54209 (accessed on 1 December 2021). [CrossRef] [Green Version]
  53. Srinivasa, M.R.; Rammohan, Y.S.; Karthik, U.N.; Sadashiva, A.N.D. Analysis on wear properties of graphene hydroxyl rein-forced aluminum composites. IJAST 2020, 29, 4162. [Google Scholar]
Figure 1. Variation of the hardness of Al nanocomposite and hybrid nanocomposites.
Figure 1. Variation of the hardness of Al nanocomposite and hybrid nanocomposites.
Materials 15 00865 g001
Figure 2. Variation of specific wear rate of different samples at a load of 3 N, sliding speed of 0.1 m/s, and a sliding distance of 100 m.
Figure 2. Variation of specific wear rate of different samples at a load of 3 N, sliding speed of 0.1 m/s, and a sliding distance of 100 m.
Materials 15 00865 g002
Figure 3. Typical 2D profiles for wear tracks after the wear test at a load of 3 N, sliding speed of 0.1 m/s and a sliding distance of 100 m: (a) Pure Al; (b) Al-0.25 wt% GO; (c) Al-10 Vol% Al2O3; (d) Al-10 Vol% Al2O3-0.25 wt% GO; (e) Al-10 Vol% Al2O3-0.5 wt% GO; (f) Al-10 Vol% Al2O3-1 wt% GO.
Figure 3. Typical 2D profiles for wear tracks after the wear test at a load of 3 N, sliding speed of 0.1 m/s and a sliding distance of 100 m: (a) Pure Al; (b) Al-0.25 wt% GO; (c) Al-10 Vol% Al2O3; (d) Al-10 Vol% Al2O3-0.25 wt% GO; (e) Al-10 Vol% Al2O3-0.5 wt% GO; (f) Al-10 Vol% Al2O3-1 wt% GO.
Materials 15 00865 g003
Figure 4. Typical frictional graphs for all the tested samples at a load of 3 N, sliding speed of 0.1 m/s, and a sliding distance of 100 m.
Figure 4. Typical frictional graphs for all the tested samples at a load of 3 N, sliding speed of 0.1 m/s, and a sliding distance of 100 m.
Materials 15 00865 g004
Figure 5. Variation of average coefficient of friction for all the tested samples at a load of 3 N, sliding speed of 0.1 m/s, and a sliding distance of 100 m.
Figure 5. Variation of average coefficient of friction for all the tested samples at a load of 3 N, sliding speed of 0.1 m/s, and a sliding distance of 100 m.
Materials 15 00865 g005
Figure 6. Typical SEM images and EDS mapping for the wear tracks of (a) Pure Al (b) Al-0.25% wt% GO (c) Al-10 Vol% Al2O3 (d) Al-10 Vol% Al2O3-0.25 wt% GO (e) Al-10 Vol% Al2O3-0.5 wt% GO (f) Al-10 Vol% Al2O3-1 wt% GO, after the wear test at a load of 3 N, sliding speed of 0.1 m/s, and a sliding distance of 100 m.
Figure 6. Typical SEM images and EDS mapping for the wear tracks of (a) Pure Al (b) Al-0.25% wt% GO (c) Al-10 Vol% Al2O3 (d) Al-10 Vol% Al2O3-0.25 wt% GO (e) Al-10 Vol% Al2O3-0.5 wt% GO (f) Al-10 Vol% Al2O3-1 wt% GO, after the wear test at a load of 3 N, sliding speed of 0.1 m/s, and a sliding distance of 100 m.
Materials 15 00865 g006
Figure 7. Typical optical counterface ball images sliding against: (a) Pure Al (b) Al-0.25 wt% GO (c) Al-10 Vol% Al2O3 (d) Al-10 Vol% Al2O3-0.25 wt% GO (e) Al-10 Vol% Al2O3-0.5 wt% GO (f) Al-10 Vol% Al2O3-1 wt% GO, after the wear at a load of 3 N, sliding speed of 0.1 m/s, and a sliding distance of 100 m.
Figure 7. Typical optical counterface ball images sliding against: (a) Pure Al (b) Al-0.25 wt% GO (c) Al-10 Vol% Al2O3 (d) Al-10 Vol% Al2O3-0.25 wt% GO (e) Al-10 Vol% Al2O3-0.5 wt% GO (f) Al-10 Vol% Al2O3-1 wt% GO, after the wear at a load of 3 N, sliding speed of 0.1 m/s, and a sliding distance of 100 m.
Materials 15 00865 g007
Table 1. Spark plasma sintering parameters used for fabricating the hybrid composite samples.
Table 1. Spark plasma sintering parameters used for fabricating the hybrid composite samples.
Die MaterialGraphite
Die diameter20 mm
Temperature550 °C
Heating rate200 °C/min
Holding time (t)10 min
Pressure (Pr.)50 MPa
Table 2. Densification of all the sintered samples.
Table 2. Densification of all the sintered samples.
SampleRelative Density (%)
Al99.8
Al-10 Vol % Al2O399.5
Al-10 Vol % Al2O3-0.25 wt% GO98.9
Al-10 Vol % Al2O3-0.5 wt% GO98.6
Al-10 Vol % Al2O3-1 wt% GO97.4
Table 3. Summary of the results obtained in the current study.
Table 3. Summary of the results obtained in the current study.
PropertySample
AlAl + 0.25 wt% GOAl + 10 Vol% Al2O3Al + 10 Vol% Al2O3+ 0.25 wt% GOAl + 10 Vol% Al2O3+ 0.5 wt% GOAl + 10 Vol% Al2O3+1 wt% GO
Hardness (HV)32.835.7655.863.5656.956
Relative density (%)99.899.299.598.998.697.4
Specific Wear Rate (10−6 mm3/Nm)5664393032518371920
Reduction in Wear Rate (%)022.6346.3755.65−47.87−238.8
Table 4. Comparative analysis of the results reports in the literature related to Al-based composites.
Table 4. Comparative analysis of the results reports in the literature related to Al-based composites.
CompositeFabrication MethodRelative Density (%)HardnessWearReduction in Wear Rate %Ref
Al-6082Stir Casting98.374 BHN62 µg0[45]
Al-6082 + 10% Al2O397.578 BHN45 µg27.41%
Al-6082 + 15% Al2O397.381 BHN35 µg43.35%
Al-6082 + 20% Al2O396.987 BHN30 µg51.61%
Al 6061Stir Casting−−95 HV0.035 gm0[46]
Al 6061 + 6% Al2O3−−105 HV0.028 gm20%
Al 6061 + 9% Al2O3−−150 HV0.023 gm34.28%
Al-6061 + 12% Al2O3−−188 HV 0.020 gm42.85%
Al-6061Stir Casting−−28 BHN1.4 × 10−3 (mm3/m)0[47]
Al-6061-5% Al2O3−−33 BHN0.9 × 10−3 (mm3/m)35.71%
Al-6061-10% Al2O3−−38 BHN0.64 × 10−3 (mm3/m)54.42%
Al-6061-15% Al2O3−−39 BHN0.66 × 10−3 (mm3/m)52.85%
AlPowder Metallurgy98111 HV0.006 (g)0[48]
Al + 0.1 wt% GNP98.898 HV0.005 (g)16.66%
Al + 1 wt% GNP98.897 HV0.007 (g)−16.66%
Al 7075Spark Plasma Sintering9996.8 HV0.0034 (mm3/m)0[49]
Al-7075/GNPs99124.9 HV0.00275 (mm3/m)19.11%
A356 Stir and squeeze casting−−−−44 Vol loss (mm3)0[50]
A356-0.5% Al2O3−−−−27 Vol loss (mm3)38.63%
A356-1% Al2O3−−−−25 Vol loss (mm3)43.18%
A356-1.5% Al2O3−−−−32 Vol loss (mm3)27.27%
AlSintering Process−−53 HRB0.039 (g)0[51]
Al + 0.01 wt%Greaphene−−54 HRB0.037 (g)5.12%
Al + 0.5 wt%Greaphene−−62 HRB0.021 (g)46.15
Al + 1 wt%Greaphene−−58 HRB0.055 (g)−41.02%
Al + 2 wt%Greaphene−−49 HRB0.105 (g)−169.23%
Al + 5 wt%Greaphene−−45 HRB0.170 (g)−335.89%
AA 6061 Microwave sintering−−65 HV0.075 (gms)0[52]
AA 6061 + 0.3%Graphene−−77 HV0.062 (gms)17.33%
AA 6061 + 0.6%Graphene−−75 HV0.062 (gms)17.33%
AA 6061 + 0.9%Graphene−−73 HV0.066 (gms)12%
AA 6061 + 1.2%Graphene−−71 HV0.070 (gms)6.66%
Al-2024Stir Casting−−−−0.109 (g)0[53]
Al-2024 + 0.25 wt% Graphene−−−−0.104 (g)4.58%
Al-2024 + 0.50 wt% Graphene−−−−0.0998 (g)8.44%
Al-2024 + 0.75 wt% Graphene−−−−0.0942 (g)13.57%
Al-2024 + 1 wt% Graphene−−−−0.0892 (g)18.16%
Al-0.25wt% GO-10Vol% Al2O3Spark Plasma Sintering98.963.56 HV251 × 10−6 mm3/Nm55.65%Present Study
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mohammed, A.S.; Aljebreen, O.S.; Hakeem, A.S.; Laoui, T.; Patel, F.; Ali Baig, M.M. Tribological Behavior of Aluminum Hybrid Nanocomposites Reinforced with Alumina and Graphene Oxide. Materials 2022, 15, 865. https://doi.org/10.3390/ma15030865

AMA Style

Mohammed AS, Aljebreen OS, Hakeem AS, Laoui T, Patel F, Ali Baig MM. Tribological Behavior of Aluminum Hybrid Nanocomposites Reinforced with Alumina and Graphene Oxide. Materials. 2022; 15(3):865. https://doi.org/10.3390/ma15030865

Chicago/Turabian Style

Mohammed, Abdul Samad, Omar Saad Aljebreen, Abbas Saeed Hakeem, Tahar Laoui, Faheemuddin Patel, and Mirza Murtuza Ali Baig. 2022. "Tribological Behavior of Aluminum Hybrid Nanocomposites Reinforced with Alumina and Graphene Oxide" Materials 15, no. 3: 865. https://doi.org/10.3390/ma15030865

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