Next Article in Journal
Improving the Interfacial Bonding Strength of Laser Direct Joining between Dissimilar 304 Stainless Steel and PCCF30 Plastic
Previous Article in Journal
Evolutions on Microstructure and Impact Toughness of G115 Steel after Long-Term Aging at 700 °C
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Heat Treatment on Microstructures and Mechanical Properties of Graphene/Aluminum Composites: Insights from Atomic Simulations

Faculty of Civil Engineering and Mechanics, Jiangsu University, Zhenjiang 212013, China
*
Authors to whom correspondence should be addressed.
Metals 2024, 14(8), 923; https://doi.org/10.3390/met14080923
Submission received: 19 June 2024 / Revised: 10 August 2024 / Accepted: 12 August 2024 / Published: 15 August 2024
(This article belongs to the Section Computation and Simulation on Metals)

Abstract

:
Graphene nanoplatelets (GNPs) have significant potential as reinforcement agents in metallic materials. Understanding the interfacial structure between GNPs and an aluminum matrix (AlM) is crucial for this composite (GNPs/AlMC) design and application. Studies on the effects of heat treatment on the interfacial character and its influence on mechanical properties at the atomic scale are, therefore, of high importance. This study presents molecular dynamics simulations investigating the effects of heat treatment on the interfacial structural evolution between GNPs and AlM and how such behavior tunes mechanical performance. The results show that the number of interfacial bonds increases significantly when the heat treatment temperature reaches 900 K, below which it stays at low level. Compared with the heat treatment temperature, the pressure has less effect on the interfacial bonds of GNPs/AlMCs. Moreover, defective GNPs with pores at certain heat treatment temperatures promote the formation of Al-C bonds at the interface and improve the mechanical properties of GNPs/AlMCs. These findings underscore the potential for utilizing heat treatment and defect engineering to improve interfacial bonding and achieve improved mechanical performance in GNPs/AlMCs.

1. Introduction

Carbon nanomaterials (carbon nanotubes, graphene, etc.) have excellent mechanical, electrical as well as optical properties [1,2,3,4,5], and are considered to be revolutionary materials of the future with important applications in materials science, micro- and nano-processing, biomedicine, the automotive industry, and aerospace industry [6,7,8,9,10,11]. Although a great deal of research has been conducted, there is still a long way to go before they can be used in practice, particularly in composite materials. For example, as the size of nanomaterials increases, a large number of defects inevitably arise, leading to a dramatic decrease in their mechanical properties [12]. Meanwhile, how to achieve high efficient load transfer remains insufficiently explored, although the incorporation of carbon nanotubes or graphene nanoplatelets (GNPs) as a reinforcing phase into the matrix to form composite materials has been considered as one of the ideal approaches to achieve their functions.
A large number of carbon nanotubes or GNP-reinforced composites have been developed, including resin matrix composites and metal matrix composites (MMCs) [13,14,15,16]. Among various MMCs, aluminum matrix composites (AlMCs) are highly favored in industries such as automotive, aerospace, and others due to their lightweight, high strength, excellent electrical, and thermal conductivity properties [17,18]. Specifically, GNP-reinforced aluminum matrix composites (GNPs/AlMCs) have received significant attention for their exceptionally high specific strength, specific stiffness, and corrosion resistance [19,20,21], and thereby hold great application potential in the aerospace industry, automobile industry, and other fields [21,22,23]. The dimensions of these components range from tens of millimeters to hundreds of millimeters. A study by Wang et al. [24] successfully fabricated GNPs/AlMCs using a flake-based powder metallurgy approach. The addition of only 0.3% GNPs in volume fraction resulted in a remarkable tensile strength of 249 MPa for GNPs/AlMCs, which was significantly improved compared to the unreinforced AlM. Xie et al. [25] proposed a deformation-driven metallurgy method for transforming powder into bulk material. By adding 1.5% GNPs in weight to AlM and achieving a uniform dispersion, the strength of GNPs/AlMCs was significantly improved, while the ductility was almost unaffected. In fact, due to the poor wettability between GNPs and AlM which leads to poor interfacial bonding strength, the enhancement of mechanical properties of GNPs/AlMCs is always limited [12,26]. Therefore, further improving the wettability and interfacial bonding strength between GNPs and AlM will be key to improving their mechanical properties.
Various interface engineering methods have been developed to enhance the mechanical properties of GNPs/AlMCs [27,28,29,30]. Guo et al. [31] found that carbide Al4C3 was formed on the interface in the process of heat treatment of GNPs/AlMCs, with the quantity and size of the carbide being highly dependent on the heat treatment temperature. It was found that the presence of carbides significantly enhanced the interfacial bonding strength of GNPs/AlMCs. It is worth mentioning the role of Al4C3 is debated in the literature. According to Guo et al., the size of Al4C3 plays a key role in modifying mechanical properties of the composites, once the size reaches microscale, it deteriorates the mechanical properties. Moreover, Al4C3 is highly hygroscopic and can absorb moisture and form acetelyne, which will have a negative impact on the mechanical properties of the composites [32]. Li et al. [33] discovered a chemical reaction occurring at the GNPs/AlMC interface as the annealing temperature increased. This reaction resulted in the formation of interface products, which improved the bonding between GNPs and AlM. And, while the formation of Al4C3 improved the mechanical properties of this composite such as hardness and wear resistance, it could also lead to nucleation of cracks and compromise overall mechanical performance. Note that once thermally treated, the formatted interfacial structures could be very rich, so they are not limited to the formation of Al4C3 [34]. Therefore, adjusting the heat treatment parameters to induce the interfacial reaction and generate an appropriate number of interfacial products can become a potential way to tune the interfacial bonding of GNPs/AlMCs and improve its mechanical properties.
In both preparation and application processes, a high-temperature environment is inevitable for the GNPs/AlMCs. To uncover how such conditions impact the structural dynamics and functionalities is essential. In this work, we use molecular dynamics (MD) simulations to investigate heat treatment of GNPs/AlMCs at the atomic scale and study how heat treatment modifies the microstructure between GNPs and AlM and ultimately tunes the mechanical properties of GNPs/MMC.

2. Simulation Methods

2.1. Atomistic Models and Molecular Dynamics Simulations

In this study, a large-scale atomic/molecular parallel simulator (LAMMPS) [35] was used to simulate the heat treatment process of GNPs/AlMCs, and OVITO (Open Visualization Tool, Basic 3.6.0) [36] software was used to visualize the interface structure dynamics of this material. Due to the fact that interfacial bonding formation and load transfer in the composite occur at the atomistic scales, the models constructed here were rational in uncovering key mechanisms; previous studies have demonstrated this in many situations. For example, when Rong et al. [37] used MD to study the mechanical properties of GNPs/AlMCs, the calculated Young’s modulus of pure aluminum was 74.78 GPa, which was very close to the experimental results. In this study, a representative model of 93 × 73 × 24 Å3 was used, a vacuum space was dug inside the aluminum block and a 68 × 50 Å2 rectangular pristine GNPs was embedded (as shown in Figure 1). The whole system contained 10,098 atoms and the volume fraction of pristine GNPs was about 7.5%. In addition, to investigate the impact of GNPs defect concentration on the interface structure of heat-treated GNPs/AlMCs, defective GNPs with two pore structures were created by selectively removing atoms from different regions of the pristine GNPs mentioned above (as shown in Figure 1b,c, named as type-A and type-B defects). In comparison to the type-A pore defects, there were more dangling bonds in the type-B defect structure. As the equilibrium bond length between aluminum and carbon atoms ranged from 2.31 to 3.36 Å [38], the distance between carbon atoms and the aluminum atoms was arranged to be greater than 3.36 Å when constructing all initial models to ensure that no chemical bonds were formed between the GNPs and AlM at the beginning. All the subsequent statistical interface bonds were formed due to heat treatment.
The entire heat treatment process includes several stages as follows: relaxation, heating, relaxation at high temperatures, cooling, and re-relaxation at a low temperature. In the relaxation process, the steepest descent algorithm was used to minimize the energy of the model to obtain the optimal geometric model. Then, the conjugate gradient algorithm and the canonical (NVT) ensemble were used for relaxation. After a relaxation period of 5 × 10−11 s at room temperature T0 (300 K), the samples were heated to the target temperature T at a rate of 2 × 1013 K/s during the heating stage. T is the temperature of heat treatment, ranging from 500 K to 900 K. Each model was maintained at T for 8 × 10−10 s during the insulation process until the structure stabilized. In the cooling process, the samples were cooled down to T0 at the same rate as the heating process. Finally, the re-relaxation of 1 × 10−10 s was applied to reach a stable equilibrium state. Three different pressure values of 1 atm, 2000 atm, and 4000 atm were applied in the whole heat treatment process, which were all feasible in experimental tests [39,40]. The heat treatment simulations from heating to re-relaxation were conducted under an isothermal and isobaric (NPT) ensemble using a Nose–Hoover thermostat, with periodic boundary conditions and a time step of 2.5 × 10−16 s.
All samples after completing heat treatment maintained the same rectangular block shape as the initial model, and subsequent mechanical property tests were performed. The x, y, and z directions had periodic boundary conditions controlled by an NPT ensemble to maintain atmospheric pressure. The time step was set at 0.25 fs.

2.2. Interatomic Potential

The ReaxFF potential [41,42], which is developed based on the covalent bonding level and is suitable for systems containing both metal and non-metal bonds, describes chemical reactions by accounting for the changes in bonding status during bond breaking and formation processes. Hong et al. [43] optimized the ReaxFF repulsive force field they developed based on a quantum mechanics (QM) training set and validated it using QM data and additional experimental data. Their results demonstrate that this interatomic potential can accurately reproduce the Al-C interaction energy obtained from QM calculations. Furthermore, this potential is applicable to high-temperature conditions, meeting the temperature requirements for the simulation of heat treatment of GNPs/AlMC in this study. The ReaxFF repulsive force field developed by Hong et al. was, therefore, employed in this study.

3. Results and Discussion

3.1. Role of Heat Treatment Temperature

This section will discuss the effect of heat treatment temperature on the microstructure and mechanical properties of GNPs/AlMCs. The microstructure of the GNPs/AlMC was significantly changed after undergoing complete heat treatment. A large number of Al-C bonds were observed at the interface between GNPs and AlM (as shown in Figure 2). In order to understand the effect of heat treatment on the interfacial bonding, R was defined in this study: R = Nb/A. Here, Nb represents the number of formed Al-C bonds in the interfacial after heat treatment, and A represents the surface area of GNPs. The obtained results are shown in Figure 3a. Each data point was calculated three times and corresponding error bars are given in Figure 3. It can be seen that the deviation of MD calculation results was small and the repeatability of the results was acceptable. When the heat treatment temperature T ≤ 800 K, there was a small and relatively low change in the R. This suggests that at low heat treatment temperatures, the formation of Al-C bonds remains limited. As the heat treatment temperature reaches 900 K, R sharply increases to 0.290. This above result indicates that the interface microstructure of GNPs/AlMC is significantly influenced by heat treatment temperature.
In order to study the effect of heat treatment on the mechanical properties of GNPs/AlMCs, uniaxial tensile simulations by MD were performed on the samples after heat treatment with a strain rate of 2 × 108/s in the x direction. Therefore, the shape and size of the tensile specimen were consistent with the previously established heat treatment model. The MD simulation results showed that, at heat treatment temperatures T ≤ 700 K, the Young’s modulus varied very slightly with temperature. Then, there was a sudden increase at 800 K, followed by a slow decrease as the temperature continued to increase (see Figure 3b). In the temperature range of 500 K ≤ T ≤ 800 K, the tensile strength and ultimate strain of GNPs/AlMC exhibited an increasing trend with higher heat treatment temperature, reaching the maximum at 800 K. By analyzing the atomic structure (as shown in Figure 4), it was observed that the content of face-centered cubic crystal structure (FCC) in GNPs/AlMCs was relatively high after lower temperature heat treatment (500 K ≤ T ≤ 800 K), which indicates that AlM maintains good crystal order and low dislocation density. The presence of numerous Al-C bonds at the interface closely interconnected GNPs with AlM, effectively fixing GNPs within AlM. This bonding mechanism facilitated the transfer of load from AlM to GNPs, thereby enhancing the mechanical properties of GNPs/AlMCs. Therefore, ordered crystal structure, some dislocations and an appropriate level of interface bonding collectively contributed to the improvement of the mechanical characteristics of GNPs/AlMCs. Computational analysis demonstrated that in the heat treatment temperature range of 500 K ≤ T ≤ 800 K, an increased number of interface bondings improved the tensile strength, Young’s modulus, and ductility of GNPs/AlMCs in general. Within the investigated temperature range, the FCC content in GNPs/AlMCs after heat treatment at 800 K reached as high as 67.6%, yielding some dislocations and impacted the overall high mechanical performance.
When the heat treatment temperature increased to 900 K, abundant dislocations occurred in the GNPs/AlMC composite (as depicted in Figure 4e). This would theoretically increase the material’s strength because of dislocation reinforcement. But, we found that this was not always the case, and several factors could possibly contribute to this observation. Firstly, high-temperature treatment leads to major changes in the crystal structure of GNPs/AlMCs; the content of the face-centered cubic (FCC) structure was only 29.0% after 900 K heat treatment. A large number of FCC structures became amorphous or partially HCP phases; such an observation was reported by Mahata et al. [44] as well, and the underlying mechanisms deserves in-depth investigations at the first principles level. The structural order of AlM was significantly reduced, which had a negative impact on the mechanical properties of GNPs/AlMCs. Secondly, the pristine GNPs were aligned parallel to the x-axis in the armchair direction. After the heat treatment at 900 K, the GNPs underwent a positional shift within the AlM, forming a certain angle with the x-axis. And, the angle and position variations were random. It not only weakened the hindering effect of GNPs on dislocation during strain, leading to further propagation of dislocation, but also elevated the challenge of load transfer to some extent. The disorder of the crystal structure and misposition and sliding of GNPs negatively affected the mechanical properties of GNPs/AlMCs. In addition, factors such as the content of interfacial bonds and crystal defects can affect the mechanical properties of GNPs/AlMCs. In future research, a combination of experimental and MD studies will be beneficial in observing detailed interface products under heat treatment pressure, as well as the mechanism responsible for the mechanical properties of GNPs/AlMC modifications.
In order to evaluate the enhancement effect of GNPs on GNPs/AlMCs, uniaxial tensile simulation of pure aluminum at room temperature was also conducted. The tensile strength, Young’s modulus, and ultimate strain of pure aluminum measured 3.3 GPa, 36.0 GPa, and 24.2%, respectively. The calculated tensile strength was higher than the experimental measurements, while the Young’s modulus was smaller, which was close to previous simulation results [37,45]. The reason can be attributed to the choice of inter-atomic potentials, defect, and size effects etc.; however, the ReaxFF used here can qualitatively well describe the atomic dynamics and overall trends [43]. Compared with pure Al, the mechanical properties of GNPs/AlMCs containing defect-free GNPs were significantly improved, and the specific reinforcement ratio is shown in Table 1. Compared with GNPs/AlMCs without heat treatment, the tensile strength, Young‘s modulus, and ultimate strain of GNPs/AlMCs heat-treated at 800 K under atmospheric pressure were increased by 22.3%, 75.6%, and 18.1%, respectively. These results show that the addition of small amounts of GNPs effectively improved the mechanical properties of AlM, and heat treatment temperatures have a significant impact on the mechanical properties of GNPs/AlMC.
Revealing the microscopic mechanism of the effect of heat treatment on the mechanical properties of materials is important for the design of GNPs/AlMCs with better performance. However, it is difficult to accurately determine the interfacial products only by MD calculations. Therefore, in order to further study the effect of bonding mechanisms on the mechanical properties of GNPs/AlMCs, it may be necessary to combine experiments.

3.2. Effect of Heat Treatment Pressure

The influence of heat treatment pressure on the microstructure and mechanical properties of GNPs/AlMCs will be discussed in this section. Shown in Figure 5a are the R values for the composite under varying pressure and temperature. The three curves are basically in the same shape, indicating that pressure had a relatively minor impact on the quantity of interfacial bonds compared to heat treatment temperature.
Under uniaxial tensile simulations, as shown in Figure 5b,c, apparent fluctuations in the mechanical properties of GNPs/AlMCs are observed. This suggests that, different from the previous section where GNPs/AlMCs were only influenced by the heat treatment temperature, the composites in this section were influenced by both the heat treatment temperature and pressure, resulting in more complex mechanical properties without clear trends. Based on the results in Figure 5b, the corresponding optimal tensile strengths of GNPs/AlMCs were observed after heat treatment at 800 K, 700 K, and 600 K under 1 atm, 2000 atm, and 4000 atm examined in this study. Furthermore, at 2000 atm, GNPs/AlMCs heat-treated within the range of 700 K to 900 K exhibited a high Young’s modulus (see Figure 5c).

3.3. Impact of Defective GNPs

Next, the effect of GNP defects on the microstructure and mechanical properties of GNPs/AlMCs were investigated. After heat treatment of GNPs/AlMCs containing defective GNPs at different temperatures, the corresponding R and mechanical properties parameters were measured, as shown in Figure 6. For comparison, the relevant data of GNPs/AlMCs containing defect-free GNPs are also presented. Two different pore structures are considered, namely type-A pores and type-B pores, as mentioned above. After heat treatment at atmospheric pressure, a significant number of interface bonds were formed at the GNPs/AlMCs interface containing defective GNPs, leading to a substantial change in the interface structure. When the heat treatment temperature T ≤ 800 K, there was little variation in the R among different samples and the value remained relatively low. When the heat treatment temperature reached 900 K, the R increased dramatically (as shown in Figure 6a,d).
The mechanical properties of GNPs/AlMC with defective GNPs also exhibited significant fluctuations. These fluctuations may be attributed to the complex micro-interface structure evolution between GNPs and AlM during heat treatment. Therefore, the internal strengthening mechanism of GNPs/AlMC becomes complicated and makes it difficult to determine the defect concentration in GNPs that yields the optimum mechanical properties. It is certain that defects in GNPs play a crucial role in the mechanical properties of GNPs/AlMCs. When the heat treatment temperature was 900 K, the composites containing 1.13% type-A defective GNPs (GNPs/[email protected]%A) exhibited improved R, tensile strength, and Young’s modulus compared to the composites containing defect-free GNPs, as shown in Figure 6a, Figure 6b and Figure 6c, respectively. The increase ratios were 10.8%, 26.9%, and 4.4%, respectively. Nevertheless, an excessive number of defects will significantly deteriorate the mechanical properties of GNPs/AlMCs. For instance, after heat treatment at 800 K under atmospheric pressure, the tensile strength and Young’s modulus of GNPs/[email protected]%A was 29.4% and 27.6% lower than those of composites containing defect-free GNPs, respectively. The Young‘s modulus of GNPs/[email protected]%A decreased more after heat treatment at 900 K. This suggests that only introducing an appropriate number of defects may be beneficial in improving the mechanical properties of GNPs/AlMCs.
The comparison of GNPs/AlMCs containing different pore defect structures was used to explore the effects of defective GNPs on the R and mechanical properties of GNPs/AlMCs. For instance, compared to GNPs/[email protected]%A, GNPs/[email protected]%B exhibited a higher R after heat treatment in the range of 500 K ≤ T ≤ 900 K. This suggests that, in the case of relatively high concentrations of defects, type-B pore defects containing more dangling bonds are prone to forming more interface bonds with AlM. After heat treatment at 700–900 K under atmospheric pressure, GNPs/[email protected]%B demonstrated better mechanical properties compared to GNPs/[email protected]%A. Particularly after heat treatment at 700 K under atmospheric pressure, the tensile strength and Young’s modulus of GNPs/[email protected]%B increased by 23.9% and 10.3%, respectively, and both exceeded the mechanical properties of GNPs/AlMC containing defect-free GNPs (see Figure 6e,f). This study indicates that the concentration and structure of GNPs defects all have an important influence on the mechanical properties of GNPs/AlMCs. And, the introduction of defects in GNPs provides a possible strategy for improving the mechanical properties of GNPs/AlMC.
It should be pointed out that this study only discussed the effect of microstructure on the mechanical properties of GNPs/AlMCs from Al-C bonds. Using MD makes it difficult to accurately obtain the number and distribution of interfacial compound products (e.g., Al4C3), which have an important influence on the interfacial strength and mechanical properties of GNPs/AlMC. Therefore, it is not comprehensive enough to consider only the influence of Al-C bonds on mechanical properties. In order to reveal the effect of interface products on the strength of GNPs/AlMC interface, further experimental research is needed.

4. Conclusions

The heat treatment process at 500–900 K of GNPs/AlMCs containing different defect concentrations of graphene under different pressures was simulated employing molecular dynamics. This is an attempt to focus on how heat treatment affects the interfacial microstructure and mechanical properties of GNPs/AlMCs at the atomic scale. The results showed that heat treatment and defect engineering of GNPs are viable choices to improve the mechanical properties of GNPs/AlMCs. The main conclusions are summarized as follows:
  • The heat treatment temperature and pressure have notable effects on the microstructure (crystal structure, dislocation, position and orientation of GNPs, interface bondings) and mechanical properties of GNPs/AlMCs. The results showed that the number of interfacial bonds increased significantly when the heat treatment temperature reached 900 K, below which it stays at a low level. In the heat treatment temperature range of 500 K ≤ T ≤ 800 K, an increased number of interface bondings improved the tensile strength, Young’s modulus, and ductility of GNPs/AlMCs in general.
  • The concentration and structure of GNPs defects play interesting roles in modifying the mechanical properties of GNPs/AlMCs via the formation of bonding structures at defect sites without a notable loss of intrinsic strength of GNP. This offers a plausible strategy for improving the mechanical properties of GNPs/AlMCs.

Author Contributions

Conceptualization, L.C., J.X., R.W. and C.T.; data curation, Z.S.; formal analysis, Z.S. and L.C.: funding acquisition, L.W. and C.T.; investigation, Z.S.; methodology, L.C., L.W., J.X., R.W. and C.T.; project administration, L.C. and C.T.; software, Z.S.; supervision, L.C. and C.T.; validation, Z.S., L.C., L.W. and C.T.; visualization, Z.S.; writing-original draft, Z.S.; writing-review and editing, L.C. and C.T. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Science Foundation of China (Grant No. 12072134, 12102151) and Postdoctoral Science Foundation of Jiangsu Province (Grant No. 2021K113B).

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors on request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Salvetat, J.P.; Bonard, J.M.; Thomson, N.H.; Kulik, A.J.; Forro, L.; Benoit, W.; Zuppiroli, L. Mechanical properties of carbon nanotubes. Appl. Phys. A 1999, 69, 255–260. [Google Scholar] [CrossRef]
  2. Papageorgiou, D.G.; Kinloch, I.A.; Young, R. Mechanical properties of graphene and graphene-based nanocomposites. Prog. Mater. Sci. 2017, 90, 75–127. [Google Scholar] [CrossRef]
  3. Zhang, Z.H.; Chen, C.F.; Guo, W.L. Magnetoelectric effect in graphene nanoribbons on substrates via electric bias control of exchange splitting. Phys. Rev. Lett. 2009, 103, 187204. [Google Scholar] [CrossRef] [PubMed]
  4. Jariwala, D.; Sangwan, V.K.; Lauhon, L.J.; Marks, T.J.; Hersam, M.C. Carbon nanomaterials for electronics, optoelectronics, photovoltaics, and sensing. Chem. Soc. Rev. 2013, 42, 2824–2860. [Google Scholar] [CrossRef] [PubMed]
  5. Zhu, G.F.; Liu, R.M.; Tang, C.; Wang, L.F. Dynamic tuning of moiré superlattice morphology by out-of-plane deformation. Appl. Phys. Lett. 2024, 124, 173508. [Google Scholar] [CrossRef]
  6. Popov, V.N. Carbon nanotubes: Properties and application. Mater. Sci. Eng. R Rep. 2004, 43, 61–102. [Google Scholar] [CrossRef]
  7. Ren, S.; Rong, P.; Yu, Q. Preparations, properties and applications of graphene in functional devices: A concise review. Ceram. Int. 2018, 44, 11940–11955. [Google Scholar] [CrossRef]
  8. Luo, S.H.; Samad, Y.A.; Chan, V.; Liao, K. Cellular graphene: Fabrication, mechanical properties, and strain-sensing applications. Matterials 2019, 1, 1148–1202. [Google Scholar] [CrossRef]
  9. Merneedi, A.; Natrayan, L.; Kaliappan, S.; Veeman, D.; Angalaeswari, S. Experimental investigation on mechanical properties of carbon nanotube-reinforced epoxy composites for automobile application. J. Nanomater. 2021, 2021, 4937059. [Google Scholar] [CrossRef]
  10. Srinivasan, V.; Kunjiappan, S.; Palanisamy, P. A brief review of carbon nanotube reinforced metal matrix composites for aerospace and defense applications. Int. Nano Lett. 2021, 11, 321–345. [Google Scholar] [CrossRef]
  11. Krasley, A.T.; Li, E.; Galeana, J.M.; Bulumulla, C.; Beyene, A.G.; Demirer, G.S. Carbon Nanomaterial Fluorescent Probes and Their Biological Applications. Chem. Rev. 2024, 124, 3085–3185. [Google Scholar] [CrossRef]
  12. Kinloch, I.A.; Suhr, J.; Lou, J.; Young, R.J.; Ajayan, P.M. Composites with carbon nanotubes and graphene: An outlook. Science 2018, 362, 547–553. [Google Scholar] [CrossRef] [PubMed]
  13. Ali, S.; Ahmad, F.; Yusoff, P.S.M.M.; Muhamad, N.; Onate, E.; Raza, M.R.; Malik, K. A review of graphene reinforced Cu matrix composites for thermal management of smart electronics. Compos. Part A Appl. Sci. Manuf. 2021, 144, 106357. [Google Scholar] [CrossRef]
  14. Moustafa, E.B.; Abdel Aziz, S.S.; Taha, M.A.; Saber, A. Influence of graphene and silver addition on aluminum’s thermal conductivity and mechanical properties produced by the powder metallurgy technique. Metals 2023, 13, 836. [Google Scholar] [CrossRef]
  15. Zhang, W.; Zhou, S.; Ren, W.; Yang, Y.; Shi, L.; Zhou, Q.; Liu, M. Uniformly dispersing GNPs for fabricating graphene-reinforced pure Ti matrix composites with enhanced strength and ductility. J. Alloys Compd. 2021, 888, 161527. [Google Scholar] [CrossRef]
  16. Wang, Y.; Cai, N.; Yang, H.; Wu, C. Application of carbon nanotubes from waste plastics as filler to epoxy resin composite. ACS Sustain. Chem. Eng. 2022, 10, 2204–2213. [Google Scholar] [CrossRef] [PubMed]
  17. Srivyas, P.D.; Charoo, M.S. Application of hybrid aluminum matrix composite in automotive industry. Mater. Today Proc. 2019, 18, 3189–3200. [Google Scholar] [CrossRef]
  18. Fereiduni, E.; Ghasemi, A.; Elbestawi, M. Selective laser melting of aluminum and titanium matrix composites: Recent progress and potential applications in the aerospace industry. Aerospace 2020, 7, 77. [Google Scholar] [CrossRef]
  19. Yu, H.; Zhang, S.Q.; Xia, J.H.; Su, Q.; Ma, B.C.; Wu, J.H.; Zhou, J.X.; Wang, X.T.; Hu, L.X. Microstructural evolution, mechanical and physical properties of graphene reinforced aluminum composites fabricated via powder metallurgy. Mater. Sci. Eng. A 2021, 802, 140669. [Google Scholar] [CrossRef]
  20. Hasan, M.S.; Wong, T.; Rohatgi, P.K.; Nosonovsky, M. Analysis of the friction and wear of graphene reinforced aluminum metal matrix composites using machine learning models. Tribol. Int. 2022, 170, 107527. [Google Scholar] [CrossRef]
  21. Alwahib, A.A.; Muttlak, W.H.; Mahdi, B.S.; Mohammed, A.Z. Corrosion resistance enhancement by laser and reduced graphene oxide-based nano-silver for 1050 aluminum alloy. Surf. Interfaces 2020, 20, 100557. [Google Scholar] [CrossRef]
  22. Shirvanimoghaddam, K.; Hamim, S.U.; Akbari, M.K.; Fakhrhoseini, S.M.; Khayyam, H.; Pakseresht, A.H.; Ghasali, E.; Zabet, M.; Munir, K.S.; Jia, S.; et al. Carbon fiber reinforced metal matrix composites: Fabrication processes and properties. Compos. Part A Appl. Sci. Manuf. 2017, 92, 70–96. [Google Scholar] [CrossRef]
  23. Zhao, Z.; Bai, P.; Du, W.; Liu, B.; Pan, D.; Das, R.; Liu, C.; Guo, Z. An overview of graphene and its derivatives reinforced metal matrix composites: Preparation, properties and applications. Carbon 2020, 170, 302–326. [Google Scholar] [CrossRef]
  24. Wang, J.; Li, Z.; Fan, G.; Pan, H.; Chen, Z.; Zhang, D. Reinforcement with graphene nanosheets in aluminum matrix composites. Scr. Mater. 2012, 66, 594–597. [Google Scholar] [CrossRef]
  25. Xie, Y.; Meng, X.; Chang, Y.; Mao, D.; Yang, Y.; Xu, Y.; Wan, L.; Huang, Y. Ameliorating strength-ductility efficiency of graphene nanoplatelet-reinforced aluminum composites via deformation-driven metallurgy. Compos. Sci. Technol. 2022, 219, 109225. [Google Scholar] [CrossRef]
  26. Sadoun, A.M.; Najjar, I.M.R.; Wagih, A. Electroless-plating of Ag nanoparticles on Al2O3 and graphene Nano sheets (GNs) for improved wettability and properties of Al–Al2O3/GNs nanocomposites. Ceram. Int. 2021, 47, 10855–10865. [Google Scholar] [CrossRef]
  27. Chu, K.; Wang, F.; Wang, X.; Li, Y.; Geng, Z.; Huang, D.; Zhang, H. Interface design of graphene/copper composites by matrix alloying with titanium. Mater. Des. 2018, 144, 290–303. [Google Scholar] [CrossRef]
  28. Zhang, X.; Shi, C.; Liu, E.; Zhao, N.; He, C. Effect of interface structure on the mechanical properties of graphene nanosheets reinforced copper matrix composites. ACS Appl. Mater. Interfaces 2018, 10, 37586–37601. [Google Scholar] [CrossRef]
  29. Liu, Z.Y.; Wang, L.H.; Zan, Y.N.; Wang, W.G.; Xiao, B.L.; Wang, D.; Wang, Q.Z.; Ni, D.R.; Ma, Z.Y. Enhancing strengthening efficiency of graphene nano-sheets in aluminum matrix composite by improving interface bonding. Compos. Part B Eng. 2020, 199, 108268. [Google Scholar] [CrossRef]
  30. Liu, P.; Xie, J.; Wang, A.; Ma, D.; Mao, Z. First-principles prediction of enhancing graphene/Al interface bonding strength by graphene doping strategy. Appl. Surf. Sci. 2020, 517, 146040. [Google Scholar] [CrossRef]
  31. Guo, B.; Chen, B.; Zhang, X.; Cen, X.; Wang, X.; Song, M.; Ni, S.; Yi, J.; Shen, T.; Du, Y. Exploring the size effects of Al4C3 on the mechanical properties and thermal behaviors of Al-based composites reinforced by SiC and carbon nanotubes. Carbon 2018, 135, 224–235. [Google Scholar] [CrossRef]
  32. Trujillo-Vázquez, E.; Pech-Canul, M.I.; Guía-Tello, J.C.; Pech-Canul, M.A. Surface chemistry modification for elimination of hydrophilic Al4C3 in B4C/Al composites. Mater. Des. 2016, 89, 94–101. [Google Scholar] [CrossRef]
  33. Li, J.; Zhang, X.; Geng, L. Effect of heat treatment on interfacial bonding and strengthening efficiency of graphene in GNP/Al composites. Compos. Part A Appl. Sci. Manuf. 2019, 121, 487–498. [Google Scholar] [CrossRef]
  34. Jiang, Y.; Xu, R.; Tan, Z.; Ji, G.; Fan, G.; Li, Z.; Xiong, D.B.; Guo, Q.; Li, Z.; Zhang, D. Interface-induced strain hardening of graphene nanosheet/aluminum composites. Carbon 2019, 146, 17–27. [Google Scholar] [CrossRef]
  35. Plimpton, S. Fast parallel algorithms for short-range molecular dynamics. J. Comput. Phys. 1995, 117, 1–19. [Google Scholar] [CrossRef]
  36. Stukowski, A. Visualization and analysis of atomistic simulation data with OVITO–the Open Visualization Tool. Model. Simul. Mater. Sci. Eng. 2009, 18, 015012. [Google Scholar] [CrossRef]
  37. Rong, Y.; He, H.P.; Zhang, L.; Li, N.; Zhu, Y.C. Molecular dynamics studies on the strengthening mechanism of Al matrix composites reinforced by graphene nanoplatelets. Comput. Mater. Sci. 2018, 153, 48–56. [Google Scholar] [CrossRef]
  38. Lee, W.; Jang, S.; Kim, M.J.; Myoung, J.M. Interfacial interactions and dispersion relations in carbon–aluminium nanocomposite systems. Nanotechnology 2008, 19, 285701. [Google Scholar] [CrossRef]
  39. Wu, L.; Liu, L.; Liu, J.; Zhang, R. Effects of high pressure heat treatment on microstructure and micro-mechanical properties of Cu77.96Al22.04 alloy. Mater. Trans. 2012, 53, 504–507. [Google Scholar] [CrossRef]
  40. Du, P.; Sun, H.; Kong, L.; Wang, Z.; Zhang, J.; Liu, W.; Xue, X.; He, Y. A study on recrystallization behavior and recrystallization texture of high pressure heat-treated Al–Mg alloy. J. Mater. Sci. 2023, 58, 2876–2892. [Google Scholar] [CrossRef]
  41. Van Duin, A.C.T.; Dasgupta, S.; Lorant, F.; Goddard, W.A. ReaxFF: A reactive force field for hydrocarbons. J. Phys. Chem. A 2001, 105, 9396–9409. [Google Scholar] [CrossRef]
  42. Senftle, T.P.; Hong, S.; Islam, M.M.; Kylasa, S.B.; Zheng, Y.; Shin, Y.K.; Junkermeier, C. The ReaxFF reactive force-field: Development, applications and future directions. npj Comput. Mater. 2016, 2, 15011. [Google Scholar] [CrossRef]
  43. Hong, S.; Van Duin, A.C.T. Atomistic-scale analysis of carbon coating and its effect on the oxidation of aluminum nanoparticles by ReaxFF-molecular dynamics simulations. J. Phys. Chem. C 2016, 120, 9464–9474. [Google Scholar] [CrossRef]
  44. Mahata, A.; Zaeem, M.A.; Baskes, M.I. Understanding homogeneous nucleation in solidification of aluminum by molecular dynamics simulations. Model. Simul. Mater. Sci. Eng. 2018, 26, 025007. [Google Scholar] [CrossRef]
  45. Zhu, Q.; Liu, X.; Yang, Q.S. Dislocation-blocking mechanism for the strengthening and toughening of laminated graphene/Al composites. Comput. Mater. Sci. 2019, 160, 72–81. [Google Scholar] [CrossRef]
Figure 1. Atomic models of (a) the constructed GNPs/AlMCs; structure of GNP in GNPs/AlMCs containing (b) type-A pores and (c) type-B pores. A color version of this figure can be viewed online.
Figure 1. Atomic models of (a) the constructed GNPs/AlMCs; structure of GNP in GNPs/AlMCs containing (b) type-A pores and (c) type-B pores. A color version of this figure can be viewed online.
Metals 14 00923 g001
Figure 2. GNPs/AlMCs with defect-free GNPs heat-treated at 500 K under atmospheric pressure of the (a) atomic structure and (b) interface bonding. A color version of this figure can be viewed online.
Figure 2. GNPs/AlMCs with defect-free GNPs heat-treated at 500 K under atmospheric pressure of the (a) atomic structure and (b) interface bonding. A color version of this figure can be viewed online.
Metals 14 00923 g002
Figure 3. Heat treatment temperature-dependent parameters of (a) R, (b) Young’s modulus, (c) tensile strength, and (d) ultimate strain for GNPs/AlMCs with defect-free GNPs heat-treated under atmospheric pressure. A color version of this figure can be viewed online.
Figure 3. Heat treatment temperature-dependent parameters of (a) R, (b) Young’s modulus, (c) tensile strength, and (d) ultimate strain for GNPs/AlMCs with defect-free GNPs heat-treated under atmospheric pressure. A color version of this figure can be viewed online.
Metals 14 00923 g003
Figure 4. Atomic structure of GNPs/AlMCs after 500–900 K heat treatment. (a) 500 K; (b) 600 K; (c) 700 K; (d) 700 K; (e) 700 K. The right subfigures introduce the type of crystal structures and dislocation represented by different colors. Light green atoms represent face-centered cubic structure, light red atoms represent close-packed hexagonal structure, light sapphire blue atoms represent body-centered cubic structure, and white atoms represent other structures. Deep blue dislocation stands for perfect dislocation, bright green dislocation stands for Shockley dislocation, rose red dislocation stands for stair-rod dislocation, yellow dislocation stands for Hirth dislocation, and light blue dislocation stands for Frank dislocation. A color version of this figure can be viewed online.
Figure 4. Atomic structure of GNPs/AlMCs after 500–900 K heat treatment. (a) 500 K; (b) 600 K; (c) 700 K; (d) 700 K; (e) 700 K. The right subfigures introduce the type of crystal structures and dislocation represented by different colors. Light green atoms represent face-centered cubic structure, light red atoms represent close-packed hexagonal structure, light sapphire blue atoms represent body-centered cubic structure, and white atoms represent other structures. Deep blue dislocation stands for perfect dislocation, bright green dislocation stands for Shockley dislocation, rose red dislocation stands for stair-rod dislocation, yellow dislocation stands for Hirth dislocation, and light blue dislocation stands for Frank dislocation. A color version of this figure can be viewed online.
Metals 14 00923 g004
Figure 5. Pressure effect on relationships of (a) R-T, (b) tensile strength-T, and (c) Young’s modulus-T, respectively, for GNPs/AlMCs with defect-free GNPs heat-treated under different pressures. A color version of this figure can be viewed online.
Figure 5. Pressure effect on relationships of (a) R-T, (b) tensile strength-T, and (c) Young’s modulus-T, respectively, for GNPs/AlMCs with defect-free GNPs heat-treated under different pressures. A color version of this figure can be viewed online.
Metals 14 00923 g005
Figure 6. Defective GNPs’ effect on relationships of (a,d) R-T, (b,e) tensile strength-T, and (c,f) Young’s modulus-T, respectively, for GNPs/AlMCs containing GNPs with type-A defects and type-B defects heat-treated under atmospheric pressure. A color version of this figure can be viewed online.
Figure 6. Defective GNPs’ effect on relationships of (a,d) R-T, (b,e) tensile strength-T, and (c,f) Young’s modulus-T, respectively, for GNPs/AlMCs containing GNPs with type-A defects and type-B defects heat-treated under atmospheric pressure. A color version of this figure can be viewed online.
Metals 14 00923 g006
Table 1. Percentage increase in mechanical parameters of GNPs/AlMCs containing defect-free GNPs at different heat treatment temperatures under atmospheric pressure compared with pure Al.
Table 1. Percentage increase in mechanical parameters of GNPs/AlMCs containing defect-free GNPs at different heat treatment temperatures under atmospheric pressure compared with pure Al.
Temperature (K)Increase in Tensile StrengthIncrease in Young’s ModulusIncrease in Ultimate Strain
50039.0%12.5%8.6%
60056.0%11.7%21.5%
70059.3%13.4%29.8%
80071.2%39.9%36.3%
90034.0%29.0%3.3%
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Shen, Z.; Chen, L.; Wang, L.; Xia, J.; Wang, R.; Tang, C. Effect of Heat Treatment on Microstructures and Mechanical Properties of Graphene/Aluminum Composites: Insights from Atomic Simulations. Metals 2024, 14, 923. https://doi.org/10.3390/met14080923

AMA Style

Shen Z, Chen L, Wang L, Xia J, Wang R, Tang C. Effect of Heat Treatment on Microstructures and Mechanical Properties of Graphene/Aluminum Composites: Insights from Atomic Simulations. Metals. 2024; 14(8):923. https://doi.org/10.3390/met14080923

Chicago/Turabian Style

Shen, Zixuan, Lei Chen, Liya Wang, Jun Xia, Ruijie Wang, and Chun Tang. 2024. "Effect of Heat Treatment on Microstructures and Mechanical Properties of Graphene/Aluminum Composites: Insights from Atomic Simulations" Metals 14, no. 8: 923. https://doi.org/10.3390/met14080923

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