Next Article in Journal
Load Frequency Control for Multi-Area Power Plants with Integrated Wind Resources
Previous Article in Journal
Use of Ecofriendly Glass Powder Concrete in Construction of Wind Farms
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Role B4C Addition on Microstructure, Mechanical, and Wear Characteristics of Al-20%Mg2Si Hybrid Metal Matrix Composite

by
Hamidreza Ghandvar
1,
Mostafa Abbas Jabbar
2,
Seyed Saeid Rahimian Koloor
3,
Michal Petrů
4,
Abdollah Bahador
5,
Tuty Asma Abu Bakar
1,* and
Katsuyoshi Kondoh
5
1
Department of Materials, Manufacturing and Industrial Engineering, Faculty of Engineering, School of Mechanical Engineering, Universiti Teknologi Malaysia, Skudai, Johor Bahru 81310, Malaysia
2
Department of Mechanical Techniques, Al-Nasiriya Technical Institute, Southern Technical University, Thi-Qar, Al-Nasiriya 64001, Iraq
3
Institute for Nanomaterials, Advanced Technologies and Innovation (CXI), Technical University of Liberec (TUL), Studentska 2, 461 17 Liberec, Czech Republic
4
Technical University of Liberec (TUL), Studentska 2, 461 17 Liberec, Czech Republic
5
JWRI, Osaka University, 11-1 Mihogaoka, Ibaraki, Osaka 567-0047, Japan
*
Author to whom correspondence should be addressed.
Appl. Sci. 2021, 11(7), 3047; https://doi.org/10.3390/app11073047
Submission received: 18 February 2021 / Revised: 22 March 2021 / Accepted: 23 March 2021 / Published: 29 March 2021
(This article belongs to the Section Materials Science and Engineering)

Abstract

:
In the current study, the effect of different B4C additions (0, 2.5, 5, and 10 wt%) on the microstructural, solidification behavior, mechanical, and tribological properties of Al-20%Mg2Si composite were studied by means of scanning electron microscopy (SEM) equipped with energy dispersive spectroscopy (EDS), X-ray diffraction (XRD), Vickers hardness, tensile, and dry sliding wear tests. The cooling curve thermal analysis (CCTA) approach was utilized to monitor the influence of B4C particles on the solidification behavior of Al-20%Mg2Si composite. The results revealed that the addition of B4C particles up to 10 wt% reduced the nucleation temperature (TN) and growth temperature (TG) of the primary Mg2Si phase. Moreover, the proper amount of B4C added to Al-20%Mg2Si composite has a significant effect on the microstructural alteration, mechanical, and tribological properties of the composite. The mean size of primary Mg2Si in Al-Mg2Si composite was 47 μm, in which with the addition of 5 wt% B4C, the particle size decreased to 33 μm. The highest UTS (217 MPa) and El% (7%) was achieved in Al-20%Mg2Si-5%B4C hybrid composite. The cast Al-20%Mg2Si composite revealed the brittle mode of fracture with some cleavage characterization, in which with the addition of 5%B4C, the fracture mode altered to a more ductile fracture. The wear results revealed that the Al-20%Mg2Si-5%B4C hybrid composite has the highest wear resistance with the lowest wear rate (0.46 mm3/Km) and friction coefficient (µ = 0.52) under 20 N applied load compared to other fabricated composites with mild abrasion as the governed wear mechanism.

1. Introduction

In recent years, the focus of materials design has moved towards lightweight, low-cost, and environmentally sustainable. According to this trend, aluminum metal matrix composites (AMMCs) have received considerable attention due to their enhanced properties (low density, brilliant cast ability, excellent mechanical properties, and good wear resistance) and low production cost [1].
Magnesium silicide, Mg2Si, displays high melting temperature (1085 °C), low density (1.99 × 103 kgm−3), high elastic modulus (120 GPa), high hardness (4.5 GPa), and a low coefficient of thermal expansion (7.5 × 10−6 K−1) [2]. As a result of the outstanding physical and mechanical properties of the Mg2Si phase, Al-Mg2Si or Mg-Mg2Si composites are desirable candidate materials for aerospace, automobile, and other applications [2]. Nevertheless, owing to low temperature brittleness, machinability, and insufficient wear resistance, the application of Al-Mg2Si composite is limited.
Among the various ceramic particles, boron carbide (B4C) particles are one of the most promising reinforcement particles as it exhibits unique physical and mechanical properties such as light-weight armor plates for its ballistic properties, wear-resistant components because of its tribological properties, and as abrasive grit. In addition, boron carbide can also be used as a neutron absorber in nuclear reactors due to the nuclear characteristics associated with its high boron content [3,4]. It has bulk modulus of about 214.8–247 MPa [5,6], low thermal conductivity of 30–42 mK, low density (2.52 g cm−3), high hardness 2900–3580 kg/mm2 and low CTE of 3.2·10−6 °C−1 [7,8]. Good chemical stability and bonding characteristics with Al alloys are other reported features of B4C particles [3]. Hence, the B4C particles are often used to reinforce Al alloys to fabricate MMCs. Baradeswaran et al. [9] identified the influence of B4C on the mechanical and tribological behavior of Al 7075 composites. They found increasing hardness of composites compared with the base alloy because of the presence of the increased ceramic phase. The wear resistance of the composites increased with increasing content of B4C particles, and the wear rate was significantly less for the composite material compared to the matrix alloy. In another study, Lashgari et al. [10] studied the influence of heat treatment on tensile properties and wear behavior of A356-10%B4C composite. They claimed that in Al-B4C composite, T6 treatment was a dominant factor on the hardness improvement in comparison with hardness increasing due to the addition of B4C hard particles. In addition, heat treatment process results in higher strength and wear properties of heat-treated metal matrix composites in comparison with an untreated state. Gudipudi et al. [11] investigated the effect of a novel ultrasonic assisted stir casting on mechanical properties of AA6061-B4C composites. Based on their findings, individual B4C distribution and the refinement of microstructure was achieved at 4 wt% B4C. The improved specific ultimate and compressive strengths at 4 wt% B4C were observed as 36.32% and 43.92%, whereas specific Vicker’s and Brinell hardness as 53.41% and 50.89% respectively. A hybrid composite contains two or more reinforcement materials; in which each of the reinforcements deliver superior properties in the composite. As a result, unique mechanical, physical, and tribological properties exist in the hybrid composites, which are not attainable in other materials. Stir casting is implemented to produce the ex-situ (hybrid) aluminium MMCs due to their simplicity and lower cost compared to other fabrication processes [12]. Moktar et al. [13] reported that with the addition of 10 wt% SiC particles to Al-20%Mg2Si composite, the morphology of primary Mg2Si was altered to polygonal and its size decreased to 30 µm. Similarly, Sukiman et al. [14] reported that when the Al-15%Mg2Si composite was added with 6 wt% YSZ, the mean size of primary Mg2Si reduced to 72.4 µm with subsequent improvement in tensile properties of the composite. Uvaraja et al. [15] examined the effect of SiC and B4C particulates-reinforced composite on Al6061 matrix. It was observed that hybrid metal matrix composites (HMMCs) showed a high level of hardness as compared to alloy because silicon carbide and boron carbide particles were uniformly embedded in the Al6061 matrix. Aziz et al. [16] studied the heat treatment and wear properties of Al2O3 and TiC particulates in reinforced Al6063 hybrid composites. The hardness of the composites was improved after heat treatment due to the presence of reinforcements in the composite, which accelerated the aging kinetics and caused an increase in dislocation density at the surrounding area of the reinforcement particles. Benal et al. [17] studied the effects of reinforcement with SiC and E-glass fibers on the wear properties of Al6061 hybrid composites. Due to the presence of both hard reinforcements and coherent precipitates in the matrix, increased hardness was observed during aging treatment. Umanath et al. [18] studied the effect of stir casting processing method to prepare Al6061-SiC-Al2O3 hybrid composites. It was observed that by adding reinforcement particles to the alloy matrix, micro-hardness increases with increased reinforcement due to solid solution hardening of the matrix. Altinkok et al. [19] evaluated the tensile behavior and microstructure of Al2O3/SiCp-reinforced AMMCs fabricated by the stir casting technique. Strength of the hybrid composite was higher than the unreinforced alloy. The significant enhancement in tensile strength of the composites was attributed to the grain and particle size, reinforcement shape, and uniform particle distribution in the alloy matrix. Uvaraja et al. [20] focused on the tribological behavior of novel hybrid composite materials by varying SiC reinforcements and a fixed percentage of B4C particles. It was observed that the increase in hardness of the composite was mainly owed to the existence of harder SiC and B4C particles in the matrix alloy, which act as obstacles to the dislocation motion. Sharma et al. [21] examined the metallography and bulk hardness of artificially aged Al6061-B4C-SiC stir cast hybrid composites. A significant improvement in hardness, tensile strength, and a marginal decrease in ductility of hybrid composites were observed compared to aluminum alloy. Uthayakumar et al. [22] studied the wear performance of Al1100-SiC-B4C hybrid composites under dry sliding conditions. An increase in hardness and wear resistance was observed due to the existence of harder SiC and B4C reinforcement particles in the matrix alloy.
Although there are several studies in the literature on the structure and properties of Al hybrid composites reinforced with B4C and other particles such as SiC, Al2O3, etc.; however, information regarding the structure and properties of Al hybrid composite reinforced with B4C and Mg2Si (Al-Mg2Si-B4C) is rather limited. Therefore, the purpose of the present study is to elucidate the influence of the different content of B4C particles on microstructure alteration, solidification behavior, mechanical properties, and wear properties of Al-20%Mg2Si composite.

2. Materials and Methods

2.1. Materials Fabrication

The initial materials to prepare Al-Mg2Si-B4C hybrid composites were pure aluminium (99.9% purity), pure magnesium (99.9% purity), pure silicon (98.9% purity), and pure B4C ceramic powder with a size of 10–30 μm. In order to fabricate the hybrid composites, Al ingot was melted at 800 °C using an induction furnace, and then Si was added to the molten metal at the same temperature. After dissolving the Si in Al, the melt temperature was reduced to 750 ± 10 °C and pure magnesium was then introduced to the melt. The temperature was reduced to prevent the oxidation of Mg. After homogenization and removal of the dross, the melt was poured into a steel mold to produce the composite ingot. To fabricate the hybrid composites, 2 kg of composite ingot was melted in a resistance furnace and then, B4C particles (2.5 wt%) were added gradually into the Al-Mg-Si melt. The melt was stirred constantly at a speed of 500 rpm for 15 min at 750 °C to obtain a uniform dispersion of reinforcement particles in the molten metal [23,24]. In fact, due to high viscosity of the composite melt, long stirring time is needed to achieve good mixing of the reinforcement particles inside the melt. The hybrid composite melt was cast in the pre-heated (220 °C) cast iron die to solidify. The procedure was repeated to prepare hybrid composites with 0, 5, and 10 wt% B4C. The fabricated samples were used for microstructure examination, mechanical tests, and wear tests.

2.2. Cooling Curve Thermal Analysis

Parallel to the casting of the composites, in order to examine the influence of B4C particles on characteristic temperatures of primary Mg2Si phase during solidification, thermal analysis was conducted, in which the molten metal was poured into a preheated cylindrical ceramic mold (800 °C for 20 min) which was embedded by a K-type thermocouple located in the middle, 15 mm above the bottom of the mold. The temperature-time data were achieved using EPAD-TH8-K and EPAD-BASE module connected to a computer with DEWESoft 7.0.5. The FlexPro 2019 data analysis software was used to extract the recorded data and smooth and plot the cooling curves, as well as the first and second derivative curves for extracting solidification parameters. For the fabricated composites, the slope of the cooling curve before nucleation temperature of primary Mg2Si was used to calculate the cooling rate. The cooling curve calculated for all thermal analysis tests was about 0.7 °C/s.

2.3. Mechanical Tests

The uniaxial tensile test was carried out based on ASTM B557M-02a by a computerized universal testing machine (Instron 5982) equipped with a strain gauge extensometer at a constant cross-head speed of 1 mm/min in room temperature. To confirm repeatability of the results, the tensile test was conducted on at least three specimens for each test. Hardness of the specimens was measured by a Matsuzawa Vickers hardness testing machine with the diamond indenter having an angle of 136° between opposite faces at a load of 5 N with a dwell of about 30 s.

2.4. Sliding Wear Test

Pin-on-disc sliding wear testing machine (TR-20, DUCOM) was utilized to investigate the wear properties of the composites according to the ASTM G99-04a standard under applied loads of 20 and 40 N, at constant parameters of 60 mm track diameter, 200 rpm sliding speed and 2000 m sliding distance, against a hardened chromium steel disc (Rc 64). The pin samples were 25 mm in length and 6 mm in diameter. The height loss of the specimens was measured using a linear variable differential transformer (LVDT) to monitor the wear loss of the specimens in microns. The volumetric loss was calculated by multiplying the height loss with cross section area of the test pin. A frictional force sensor was used to measure the generated frictional force on the specimens in Newton. A set of three samples was tested for each experimental condition and the average values were used for further analysis.

2.5. Microstructure Characterization

For microstructure characterization, the specimens were cut from the middle portion of cast billets and then underwent the grinding and polishing procedures in accordance with the standard metallography routines. Microstructures were analyzed using scanning electron microscopy, SEM (Hitachi S3400N) equipped with EDX facility. The fracture surfaces of the tensile test as well as the worn surface of the wear test specimens were cleaned with acetone using ultrasonic cleanser and examined using SEM. An I-Solution image analyzer was used to calculate the quantitative characteristics of the microstructure density, and aspect ratio as
Mean   density   =   1 m j = 1 m 1 n i = 1 n D i i
Mean   aspect   ratio   = 1 m j = 1 m 1 n i = 1 n a i b i i
where Di, ai, and bi are the density, longest, and shortest dimensions of a single Mg2Si particle respectively, n is the number of Mg2Si particles measured in a single field (0.60126 mm2), and m is the number of the evaluated fields. Furthermore, X-ray diffraction (XRD) (PHILIPS binary diffractometer with Cu-kα radiation application) was utilized to analyze the phase constituents of the fabricated hybrid composite.

3. Results and Discussion

3.1. Microstructure Examination

The microstructural micrograph of Al-20%Mg2Si-10% B4C hybrid composite is shown in Figure 1a. As seen, the composite structure consists of α-Al as matrix and Mg2Si and B4C as reinforcement particles. The corresponding elemental mapping as well as the XRD patterns of Al-20%Mg2Si-10% B4C hybrid composite is illustrated in Figure 1b and Figure 2 respectively, which indicate the existence of the Al, Mg2Si and B4C in the HMMC.
Figure 3a–d exhibits the SEM micrographs of Al-20%Mg2Si composite added with different B4C concentrations. As seen in Figure 3a, in the composite without B4C particles (base), primary Mg2Si particles are large and irregular and often in polyhedral shape surrounded with eutectic Mg2Si cells. With the addition of 2.5 wt% B4C into the Al-Mg2Si composite, the morphology of primary Mg2Si is still in polyhedral shape, as shown in Figure 3b. However, when the composite was treated with 5 wt% B4C, the morphology of primary Mg2Si slightly altered to polygonal as depicted in Figure 3c. Moreover, B4C particles are distributed homogeneously throughout the aluminium matrix. With further addition of B4C particles until 10 wt%, the Mg2Si particles retain the polygonal morphology; however, some agglomeration of B4C particles are observed in the composite area as illustrated in Figure 3d. Similar behavior of produced agglomeration as a result of increased B4C reinforcement content in Al composites was reported by other researchers [25,26,27]. This phenomenon can be attributed to the insignificant wettability of ceramic particles with aluminum melt as the particles float on top of the melt surface owing to the surface tension and high interfacial energy of the reinforcing particles [28]. Besides, the lower density of B4C particles compared to the aluminum causes the B4C particles to float and leads to the non-uniformed dispersion where the B4C particles agglomerate and segregate at a particular place in the melt.

3.2. Quantitative Analysis of Microstructure

Figure 4 shows that the alterations in average particle size, aspect ratio, and particle formation per unit area (density) correspond to the various B4C additions. The results match well with the microstructure observation in Figure 3, in which with the addition of 2.5 wt% B4C, the average particle size decreased by 13% from 47 to 41μm, the aspect ratio decreased by 3.7% from 1.34 to 1.29, and the density also decreased by 16% from 135 to 113 particle/mm2. With the addition of 5 wt% B4C to Al-20%Mg2Si composite, the particles size, aspect ratio, and density of primary Mg2Si further decreased to 33 μm, 1.27 and 94 particle/mm2. However, as observed in Figure 3, with a further addition of B4C to 10 wt%, the mean size and aspect ratio of primary Mg2Si increased to 38 µm and 1.28 respectively. Nevertheless, the density decreased to 63 particle/mm2 (Figure 4). Therefore, according to the results in Figure 3 and Figure 4, it can be observed that Al-20%Mg2Si-5%B4C hybrid composite illustrates the best condition in terms of primary Mg2Si refinement and B4C particles distribution.

3.3. Influence of B4C Addition on Cooling Curves

Figure 5 demonstrates the cooling curve of Al-20%Mg2Si composite with the first and second derivative curves corresponds to different contents of B4C. The peaks on the first and second derivative curves represent distinct phase transformations that occurred during solidification. According to the binary phase diagram of Al-20%Mg2Si, the first phase is the primary Mg2Si, while the second phase is the formation of eutectic Al-Mg2Si. It is observed that the addition of B4C altered the solidification behavior of the Al-20%Mg2Si composite by changing the solidification temperatures (nucleation temperature (TN), growth temperature (TG)) of primary Mg2Si phase that is correlated with the microstructures of the composites, as observed in Figure 3.
Figure 6 shows the nucleation temperature (TN) and growth temperature (TG) of the primary Mg2Si phase which correspond to the various additions of B4C. The plotted graph of TN has the same pattern as TG of primary Mg2Si, representing the same influence of the B4C on these temperatures. Based on the result, it can be observed that the TN of primary Mg2Si decreased when the B4C addition was introduced from 2.5 to 10 wt%. The same trend of graph is also observed for the TG of primary Mg2Si phase. From the temperature of 659.2 °C for base Al-20%Mg2Si composite, TN of primary Mg2Si decreased to 659 °C, 657 °C, and 655.8 °C after the addition of 2.5, 5, and 10 wt% B4C respectively.
Figure 7 depicts the duration (Δt) of primary Mg2Si particles formation. The tGMg2Si-tNMg2Si is used to indicate the time needed for the primary Mg2Si to nucleate and grow until complete solidification before the beginning of the eutectic cells. In the base Al-20%Mg2Si composite, the duration of primary Mg2Si to grow is about 13.82 s. The duration is reduced to 12.59 and 10.32 s as the composite was added with 2.5 and 5 wt% B4C, respectively. Nonetheless, the duration of primary Mg2Si particles formation increased to 12.42 s when the composite is treated with 10 wt% B4C. The difference between the duration is small, which is no more than 3.5 s. It is probable that B4C addition plays a role in controlling the growth of Mg2Si particles, which results in the changing particle size. The obtained results demonstrate that when Al-20%Mg2Si composite is added with 5%B4C particles, the duration of primary Mg2Si formation is less than other B4C additions, which is consistent with the results in Figure 3 and Figure 4, in which the addition of 5%B4C to Al-20%Mg2Si composite caused the most refinement effect on primary Mg2Si particles.

3.4. Refinement Mechanism of B4C Addition

The results obtained from microstructure and thermal analysis clearly demonstrate that addition of B4C particles to Al-Mg2Si composite caused partial refinement of the Mg2Si particles. The appropriate refinement has been obtained with addition of 5 wt% B4C. Figure 8 illustrates the SEM microgrpah of primary Mg2Si particles including dark particle in its center with corresponding line scanning spectra analysis. As seen, the EDS analysis reavled that the dark color particle in the centre of primary Mg2Si is B4C. The existance of this particle in the center of Mg2Si crystal propose that B4C particle may provide favorable nucleation substrates for the Mg2Si particles precipitation.
In order to find out whether the refining effect is related to the role of B4C particles as heterogeneous nucleation sites for primary Mg2Si, the crystal structures of Mg2Si and B4C were compared and the lattice misfit between these crystal structures was calculated. The crystal structures of B4C and Mg2Si are shown in Figure 9. It can be clearly observed that the atoms distributed on the B4C and Mg2Si planes are in a different manner. In addition, the calculated lattice misfit between the B4C and Mg2Si is 43.6%. According to Bramfitt’s two-dimensional lattice misfit theory, if the mismatch between two planes is below 15%, one phase can act as a heterogeneous nucleation substrate for another [29]. Thus, it is observed that B4C particles is not likely to meet the character of heterogeneous nucleation substrates for primary Mg2Si crystals.
In addition, Figure 6 depicts that with the addition of B4C particles to Al-Mg2Si composite, the nucleation temperature (TN) of primary Mg2Si was reduced. It has been reported that provision of heterogeneous nucleation substrates for primary Mg2Si particles through addition of refiners occurred at higher nucleation temperatures which results in a refined microstructure of the composites [30,31]. Hence, it can be concluded that the refinement of Mg2Si particles as a result of B4C addition is related to the growth stage rather than nucleation. The possible reason for refinement of primary Mg2Si particles is the presence of the B4C particles that act as a physical barrier during the growth of the primary Mg2Si phase during solidification. According to the particle pushing phenomenon, the B4C particles are rejected by the solidification front and finally stuck in grain boundary areas [32].
It was expected that with the addition of higher content of B4C particles (10 wt%) to Al-20%Mg2Si composite, the size of the primary Mg2Si decreases. However, as shown in Figure 3 and Figure 4, addition of 10 wt% B4C results in a reverse trend as the mean size of primary Mg2Si particles increased. In fact, in Al-20%Mg2Si-10%B4C hybrid composite, the B4C particles are agglomerated in some areas in the matrix with non-uniform distribution throughout the composite matrix; therefore, they cannot effectively hinder the growth of primary Mg2Si particles in the preferred Mg2Si growth directions, which results in an increase in the particles size. Moreover, the average number of primary Mg2Si particles per unit (mm2) (density) decreased with the increasing B4C concentration up to 10 wt%, as illustrated in Figure 3. The drop in the density of primary Mg2Si particles as a result of B4C addition can be associated with the formation of some boron-containing compounds in the matrix of the composite. According to Shirvanimoghaddam et al. [25], and Li et al. [33], some B-containing compounds are detected in the various B4C reinforced aluminium composites, in which some B elements are dissolved from B4C particles into the liquid-solid interface. As observed in Figure 10, the SEM micrograph with corresponding EDX results depicts that the presence of these boride compounds consist of Al and Mg. The formation of Al-Mg-B consumes the existed Mg in the melt, and hence, the amount of Mg atoms to react with Si to produce Mg2Si decreased. This findings are aligned with the results by Azarbarmas et al. [34] where 1%B was added to Al-15%Mg2Si composite. In fact, partial dissolution of B4C in Al leads to the formation of B-containing compounds. Viala et al. [35] have pointed out that as a result of the reaction between Al and B4C at temperatures above 700 °C, AlB2, Al3BC, and Al4BC are produced. The dissolution of B4C immediately saturates the molten Al with carbon and boron since the solubility of these elements in Al is low [36]. However, as the amount of Al-Mg-B phase is small, it is difficult to detect them in the XRD patterns (Figure 2). The observation in the density of primary Mg2Si particles as a result of increasing B4C content is consistent with the decreasing trend of nucleation temperature (TN) of primary Mg2Si phase in Figure 6. In fact, the reduction of Mg in the melt as a result of the formation of Al-Mg-B compound leads to a shift in the composition of the Al-Mg2Si composite to lower the Mg2Si content (eutectic point); hence, the nucleation temperature of primary Mg2Si phase decreased.
According to Azarbarmas et al. [34], in the Al-Mg2Si composite treated with B element, beside the segregation of boron on the solid-liquid interface, some B atoms also adsorbed on the growth front and changed the surface energy of the Mg2Si crystals. Nevertheless, in the present study, due to the interaction of B with the melt and formation of B-rich intermetallic compounds, more B atoms are consumed, in which the extra B atoms are inadequate in the melt to absorb the primary Mg2Si growth front and poison its anisotropic growth and alter its morphology significantly. Therefore, as shown in Figure 3 and Figure 4, addition of B4C particles into Al-20%Mg2Si composite has a minor impact on the aspect ratio of primary Mg2Si particles, in which the maximum alteration of the aspect ratio occurred after the addition of 5 wt% B4C (5.22%) compared to the base composite. This result demonstrates that B4C addition has no obvious changes on the structure of primary Mg2Si particles.

3.5. Mechanical Properties

3.5.1. Hardness

The average Vickers hardness of the fabricated composites as a result of different B4C additions is illustrated in Figure 11. It can be observed that with the addition of 2.5 and 5.0 wt% B4C ceramic particles to Al-20%Mg2Si composite, the hardness value increased to 79 and 100 HV, respectively in comparison with the base composite (75 HV). The enhancement in the hardness property is attributed to the refinement of primary Mg2Si particles as well as increasing the content of B4C reinforcements. As observed, the hardness value of Al-20%Mg2Si-5%B4C hybrid composite is greater than other fabricated composite, which is due to the uniform distribution of B4C particles. The existence of hard B4C and fine Mg2Si particles increased the obstacles to the grain boundary sliding [37]. Another factor for the improvement of the hardness of the composite due to B4C addition is increase in the dislocation density as a result of thermal expansion mismatch among the reinforcement and matrix at the particle-matrix interfaces [11]. These factors result in high restrictions to the localized deformation during indentation. However, with further addition of B4C particles to 10 wt%, the hardness value decreased to 84 HV. The reason for the decreasing hardness value is due to the agglomeration of B4C particles in the matrix of the composite (Figure 3d). In addition, as depicted in Figure 4, by increasing the B4C concentration to 10 wt%, the size of primary Mg2Si particles increased and the density of primary Mg2Si particles in the composite decreased as well, hence, the number of the Mg2Si particles to resist the applied load during the hardness test is reduced. Therefore, both B4C agglomeration and low density of primary Mg2Si particles cause a higher exposure of the soft aluminium matrix to the intender, and as a result, the average hardness of the composite is reduced.

3.5.2. Tensile Properties

The engineering stress-strain curves of Al-Mg2Si composite added with various contents of B4C ceramic particles is illustrated in Figure 12a. The ultimate tensile strength (UTS) and elongation percentage to failure (El%) values of the composites are demonstrated in Figure 12b,c respectively. As seen, the addition of 2.5 and 5 wt% B4C to Al-20%Mg2Si composite increased the tensile properties, in which the addition of 5 wt% B4C results is the highest value as the ultimate tensile strength (UTS) and percentage elongation (El%) increased from 177 MPa and 4.6% in the base Al-20%Mg2Si composite to 217 MPa and 7% in Al-20%Mg2Si-5%B4C hybrid composite. It can be observed that there is a consistency between these results and the microstructural observations in Figure 3 as Al-20%Mg2Si-5%B4C hybrid composite exhibits a uniform distribution of B4C reinforcements as well as refined primary Mg2Si particles. The high strength values of Al-20%Mg2Si-5%B4C hybrid composite is due to the existence of B4C particles in the composite matrix with uniform distribution, which creates dislocation pile-up in their neighborhood. Thus, the strength of the composite enhanced as a result of increasing dislocation interactions and dislocation density in the matrix near interfaces of the matrix and reinforcement.
In addition, partial refinement of primary Mg2Si particles has a significant impact on strength enhancement of the composite. As depicted in Figure 3a, in base Al-20%Mg2Si, the primary Mg2Si particles exist with polyhedral shape and sharp tips, in which these sharp areas are potential sites for stress concentration which cause a drop in tensile properties. Nevertheless, in Al-20%Mg2Si-5%B4C hybrid composite, the primary Mg2Si particles are finer and their morphology is regular (Figure 3c) compared to the base material, which results in tensile improvement. The most probable explanation of this behavior could be a deformation stregthening of hybrid composites which are brittle in nature but demonstrate higher strength and elongation when added with B4C particles due to Hall-Petch dependence on the Mg2Si particles size. The effect of particles size on tensile strength can be discribed through dislocation theory. The formation of plastic deformation needs to overcome the ultimate stress, which can attribute to a mass of moving dislocation. However, grain boundary will hinder the dislocation glide, and even leads to the dislocation pile-up. When the content of dislocation is pile-up increases a certain value, it will become the driving force of dislocation glide [37]. Furthermore, the particle size becomes larger, the content of dislocation pile-up is more; and the driving force of dislocation glide is bigger. That is why alloy with small particle size exhibits high strength. At the same time, due to small particle size increasing the number of particle boundary, the ability of hindering the dislocation motion is high. Therefore, the UTS for the plastic deformation is strongly influenced by Mg2Si particles size.
However, as shown in Figure 12b,c, when the concentration of B4C increased to 10 wt%, the UTS and El% of the composite reduced to 150 MPa and 2.7%, which is lower than other fabricated composites. The reduction in the tensile properties of the composite from the addition of higher B4C concentration (10 wt%) is due to the agglomeration of B4C particles as well as their non-uniform distribution in the matrix (Figure 3d). In fact, in the agglomerated areas, the bonding strength between the Al matrix and B4C particles is low and the load transferring from the matrix to B4C particles is not appropriate. Such insufficiencies cause a crack or decohesion of nucleation sites, which lead to composite failure at low stress levels. Furthermore, the addition of 10 wt% B4C to Al-20%Mg2Si composite results in increment in the average size of primary Mg2Si particles (Figure 3 and Figure 4), which leads to a decrease in the tensile properties. A bigger size of primary Mg2Si provides more potential substrates for nucleation and propagation of a crack, resulting in the failure of the composite by particle fracture [38].

3.5.3. Fracture Surface

The SEM images of tensile fracture surfaces of the Al-20%Mg2Si composite added with different percentages of B4C are exhibited in Figure 13a–f. Primary Mg2Si particle is intrinsically brittle and has more intrinsic defects, which cause cellular fracture mode in the fracture surface of base Al-20%Mg2Si composite due to cleavage in fracture planes of coarse Mg2Si particles that are indicated by arrows (Figure 3a). Furthermore, the sharp corners of the irregular primary Mg2Si particles in the base Al-Mg2Si composite serve as stress concentration sites for crack initiation and propagation which results in fracture of these coarse particles (Figure 3b) [39].
However, after adding B4C particles up to 5 wt% to the Al-20%Mg2Si composite, the fracture mode changed from brittle in the base composite to ductile, as depicted in Figure 13c,d. An interesting result is achieved for Al-20%Mg2Si-5%B4C hybrid composite, in which the fracture surface contains decohered primary Mg2Si particles and a more ductile area. The obtained result is because of the presence of partially refined primary Mg2Si particles and also B4C particles, which have uniform distribution and good bonding with the Al matrix. The matrix alloy can be reinforced with a uniform distribution of the particle due to the effective role of the particle-rich region in the prevention of crack propagation; however, a particle-poor region offers canals for the propagation of crack [40]. Due to the decreasing size and aspect ratio of Mg2Si particles after the addition of 5 wt% B4C particles (Figure 4), the Mg2Si provides less potential areas for crack nucleation compared to the base composite, and hence crack initiation is hindered. Furthermore, as the bonding between B4C particles with the matrix is good, fracturing almost occurs in the B4C particles rather than in the particle-matrix interface. This is in accordance with the model proposed by Babout et al. [41] in simulation of the fracture behavior in MMCs. Based on their suggestion, regardless of particle density, the fracture of interface is strength controlled, and decohesion occurred once the tensile stress in the particles exceeded the interface strength as a critical value. Based on these results, it is concluded that the Al-B4C interface is steady and stable, hence, the initiation and propagation of the cracks is from Mg2Si particles and not the Al-B4C interface. Moreover, as seen in Figure 13d, the fracture surface exhibits several dimples, which are indications of a ductile fracture. This result is aligned with high strength and elongation of composite, as shown in Figure 12.
Figure 13e,f depicts the fracture surface of Al-20%Mg2Si-10%B4C hybrid composite. As observed, B4C particles agglomerate in some areas on the fracture surface. The particle agglomeration increases the local stresses and proffers substrates for nucleation of crack and low energy propagation canals across the linked brittle particles. Failure can be portrayed by higher stresses generated in these areas. The interface of B4C/matrix where the B4C particles agglomerate shows interfacial decohesion, in which the failure took place along this interface. If the bonding strength in the interface between particle and matrix is weak, debonding occurs between matrix and particles before a particle fracture [39]. As the B4C particles debonded from the matrix in a brittle manner, this feature reveals a weak interface bonding which creates favorable substrates for the initiation of crack and drops the load bearing capacity of the composite. In this condition, the crack needs less energy to propagate through the interface which can be reconciled with low strength and elongation of composite, as depicted in Figure 12. The crack initiates at the interface between particle and matrix, then propagates through the matrix, and finally joins other cracks, resulting in composite failure.
Porousity and bifilm on fracture surface are clearly observed in Figure 13e. It has been reported in many investigations that with increasing volume fraction of the reinforcement particles, internal defects, such as porousity and oxide layer (bifilm) expand [42]. During composite fabrication via vortex system and pouring of the molten metal matrix composite in the mould cavity, the bifilms or folded-over films are entrained, leading to successive impairing of UTS and El% of the composite (Figure 12). Not only does the unbounded oxide-oxide interface of the bifilms introduce stress concentration, the films also act as pre-existing microcracks.

3.6. Wear Behavior

Figure 14 demonstrates that the wear rate of Al-20%Mg2Si composites correspond to different B4C additions under the applied loads of 20 and 40 N. As seen, with the addition of 5 wt% B4C, the wear resistance increased when the wear rate of Al-20%Mg2Si composites decreased from 0.85 mm3/km to 0.46 mm3/km under 20N applied load. This improvement was the result of the load bearing capacity of the composites by ceramic particles during sliding. This was similar to the results reported by [43] in Mg-TiO2 magnesium matrix composite, where, the addition of 5 wt% TiO2 into the Mg matrix improved the wear resistance of the composite. The enhancement in wear resistance could be associated with the increase in hardness, in which the hardness of the Al-20%Mg2Si composites with the addition of 5 wt% B4C increased to 100 HV (Figure 11). The connection between hardness and wear rate can be explained by Archard’s law, which describes an inverse proportionality between hardness and the wear rate of a material [44].
Another reason for the decrease in the wear rate of the composite is due to the partial refinement of primary Mg2Si particles, as shown in Figure 3 and Figure 4. It was expected that during the wear test, the wear rate of the hybrid composite with a finer Mg2Si size decreases because grain boundary strengthening occurs with smaller particle size, leading to strain hardening [45]. With the addition of 10 wt% B4C to the Al-20%Mg2Si composites, the wear resistant was supposed to increase; however, as seen in Figure 14, the wear rate increased to 0.73 mm3/km, causing the increase in material removal. The high wear rate of Al-20%Mg2Si-10%B4C hybrid composite was due to the agglomeration of the B4C particles in the matrix (Figure 3d), which resulted in the ‘three-body’ wear during sliding and increment of the composite’s wear rate. Additionally, due to the fragile bonding between the matrix and B4C in the agglomerated regions, the B4C particles were unable to support the applied load efficiently. Therefore, the interface offered a site for crack nucleation and tended to pull out the particles from the wear surface, which led to the formation and propagation of crack in the subsurface wear region [46]. In addition, as mentioned before in Figure 4, increasing the content of B4C to 10 wt%, decreased the density of primary Mg2Si particles. As a result, the load bearing effect of the hybrid composite reduced and during the wear rate the soft Al matrix was subjected to the asperities of the steel disc which led to an increase in the wear rate. Figure 14 also illustrates the increased wear rate when the applied load is increased from 20N to 40N. This is expected as the number of contact asperities and the separation of surface asperities increased at the higher loads. Increasing the applied load also contributed to the accumulation of wear debris in the space between the pin and disc that led to a more serious abrasive wear, increasing the wear rate.
The coefficient of friction (COF) of the Al-20%Mg2Si composites added with various content of B4C for the applied loads of 20 and 40 N is depicted in Figure 15. As seen, similar to the trend of wear rate, under the applied load of 20N, the COF (µ) of the Al-20%Mg2Si composite decreased from 0.46 in the base Al-20%Mg2Si composite to 0.43 and 0.39 when the composite was treated with 2.5 and 5 wt% B4C respectively. Hence, Al-20%Mg2Si-5%B4C hybrid composite has the lowest COF compared to other fabricated composites. The reason for the reduction of the COF in the composite added with 5% B4C can be related to the increased hardness of the composite (Figure 11) as finer Mg2Si particles can support more load, and decrease the contact area between the pin and the sliding disc. This indirectly prevents the pin surface from acquiring more scratches. In addition, Devaraju, et al. [47] reported that the segregation of boron which acted as a dry lubricant between the sliding surfaces, leads to a decreased friction coefficient.
However, with a further increase in the B4C content to 10%, the COF increases to 0.45. Due to the agglomeration of B4C particles in 10 wt% B4C hybrid composite, the minority of the Mg2Si particles appeared in coarse size. The fracture of the large and brittle Mg2Si could initiate the crack propagation in the matrix, and small Mg2Si particles are pulled out from the matrix, which further acts as the source for abrasive wear, causing the COF to increase. The increasing COF could also is due to the fragmentation of boron oxide layer. B2O3 layer occurred when the pulled out B4C particles reacted with the atmosphere. Its delamination during dry sliding formed scratches and grooves on the pin surface that increased the friction between the pin and the sliding surface [48].
An increase in the applied load to 40 N increases the COF of the composites compared to the 20 N applied load; however, the Al-20%Mg2Si composite added with 5 wt% B4C owns the lowes COF (0.52) compared to other fabricated composites and addition of 10 wt% B4C incresed the COF to 0.54, which is similar with the trend of COF of the composites under 20N.
Figure 16a–f demonstrates the worn surface of the fabricated composites added with 0, 5, and 10 wt% B4C particles under the applied loads of 20 and 40 N. The worn surface of Al-Mg2Si composite (Figure 16a) displays several continuous grooves parallel to the sliding direction. Moreover, the counterfaces between the pin and the sample surface increased; hence, more materials were detached from the sample’s surface. This detachment may be due to the adhesive friction that subsequently increased the wear rate (Figure 14). The principal wear mechanisms of the samples were abrasion and delamination wear. By increasing the load to 40 N, delamination and adhesion govern the wear mechanisms, in which extensive pits and large size of delamination flakes exist on the worn surface, indicating the alteration of wear mechanism to severe wear (Figure 16b). With the addition of 5 wt% B4C particles to the Al-20%Mg2Si composite, the delamination pit areas as well as the groove width decreased, illustrating a mild abrasive wear mechanism, as shown in Figure 16c. This is consistent with decreasing wear rate and coefficient of friction results depicted in Figure 14 and Figure 15, respectively.
On the other hand, as observed in Figure 16e, by adding 10 wt% B4C to the Al-20%Mg2Si composite, deep parallel grooves with large widths indicate a severe abrasion, which is regularly produced by rigid particles between the pin and disc that cut into the worn surface. As observed in Figure 3d, the introduction of a high percentage of B4C particles (10 wt%) to the Al-20%Mg2Si composite caused the B4C particles to cluster in some regions in the matrix with a non-uniform distribution. Therefore, during the sliding wear test, the agglomerated B4C particles were pulled out from the matrix, fragmented, and became entrapped between the pin and disc. As a result, abrasion took place, leading the material to be omitted from the abrasion groove Figure 16e. Hence, the predominant wear mechanisms were severe abrasion and delamination. In addition, by increasing the applied load to 40 N, the flake-like delamination pit areas as well as abrasion grooves become wider and deeper, as seen in Figure 16f. This extensive delamination is due to crack nucleation and propagation sourced from the fracture and detachment of B4C particles.

4. Conclusions

The effects of various B4C additions on the microstructural, mechanical, and tribological properties of Al-20%Mg2Si composite were investigated. It can be concluded that:
  • With the addition of 5 wt% B4C, the mean size of primary Mg2Si in Al-20%Mg2Si composite decreased from 47 to 33 µm, and its morphology was altered from octahedral to polygonal shape. However, with further addition of B4C to 10 wt%, the morphology of primary Mg2Si remains in polygonal shape with an increase in its size to 38 µm.
  • The tensile result showed that the best level of B4C was 5 wt% in which the UTS and El.% of Al-20%Mg2Si composite increased from 177 MPa and 4.6% to 217 MPa and 7% in the Al-20%Mg2Si-5%B4C hybrid composite, respectively. The mechanisms of strengthening of the composites were load transfer effects, dislocation strengthening, and grain refinement. In addition, the Vickers hardness value of Al-20%Mg2Si composite increased from 75 Hv to 100 Hv after the addition of 5 wt% B4C particles to the composite.
  • Examination of the fracture surface revealed that the fracture of the Al-20%Mg2Si composite was brittle (including the cleavages). With the addition of 5 wt% B4C particles, the fracture surface owned more dimples, which represents the domination of the ductile fracture.
  • The results of the sliding wear test showed that the addition of 5 wt% B4C to the Al-20%Mg2Si composite improved the wear resistance of the alloy by decreasing the wear rate by 46%. The enhancement in wear resistance was due to the self-lubrication function of B4C, uniform distribution of B4C particles and good bonding of B4C with the matrix as well as the small size of primary Mg2Si particles, which in turn decreased the friction coefficient of the composite.
  • Examination on the worn surface revealed that the governed wear mechanisms of the Al-20%Mg2Si composite were abrasion and delamination, which changed to mid abrasion in the Al-20%Mg2Si-5%B4C hybrid composite.

Author Contributions

H.G., M.A.J., and A.B. methodology, formal analysis, investigation, writing—original draft; H.G., K.K., and S.S.R.K. conceptualization, methodology, software, formal analysis, investigation, writing—original draft; T.A.A.B., K.K., and M.P. validation, resources, supervision, writing—review & editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research is financially supported by Universiti Teknologi Malaysia (UTM) under the research grants 04E95 and 06G22. In addition, this research was supported by the Ministry of Education, Youth, and Sports of the Czech Republic and the European Union (European Structural and Investment Funds Operational Program Research, Development, and Education) in the framework of the project “Modular platform for autonomous chassis of specialized electric vehicles for freight and equipment transportation”, reg. no. CZ.02.1.01/0.0/0.0/16_025/0007293.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors acknowledge Universiti Teknolgi Malaysia (UTM) for provision of research facility.

Conflicts of Interest

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

References

  1. Srivastava, A.; Dixit, A.; Tiwari, S. A Review on Fabrication and Characterization of Aluminium Metal Matrix Composite (AMMC). Int. J. Adv. 2014, 2, 516–521. [Google Scholar]
  2. Lu, L.; Lai, M.O.; Hoe, M.L. Format on of nanocrystalline Mg2Si and Mg2Si dispersion strengthened Mg-Al alloy by mechanical alloying. Nanostruct. Mater. 1998, 10, 551–563. [Google Scholar] [CrossRef]
  3. Bonnet, G.; Rohr, V.; Chen, X.-G.; Bernier, J.-L.; Chiocca, R.; Issard, H. Use of Alcan’s Al-B4C metal matrix composites as neutron absorber material in TN International’s transportation and storage casks, Packaging, Transport. Storage Secur. Radioact. Mater. 2009, 20, 98–102. [Google Scholar] [CrossRef]
  4. Pozdniakov, A.; Zolotorevskiy, V.; Barkov, R.; Lotfy, A.; Bazlov, A. Microstructure and material characterization of 6063/B4C and 1545K/B4C composites produced by two stir casting techniques for nuclear applications. J. Alloy. Compd. 2016, 664, 317–320. [Google Scholar] [CrossRef]
  5. McClellan, K.J.; Chu, F.; Roper, J.M.; Shindo, I. Room temperature single crystal elastic constants of boron carbide. J. Mater. Sci. 2001, 36, 3403–3407. [Google Scholar] [CrossRef]
  6. Dodd, S.P.; Saunders, G.A.; James, B. Temperature and pressure dependences of the elastic properties of ceramic boron carbide (B4C). J. Mater. Sci. 2002, 37, 2731–2736. [Google Scholar] [CrossRef]
  7. Pozdniakov, A.; Lotfy, A.; Qadir, A.; Shalaby, E.; Khomutov, M.; Churyumov, A.; Zolotorevskiy, V. Development of Al-5Cu/B4C composites with low coefficient of thermal expansion for automotive application. Mater. Sci. Eng. A 2017, 688, 1–8. [Google Scholar] [CrossRef]
  8. Canakci, A.; Arslan, F.; Varol, T. Effect of volume fraction and size of B4C particles on production and microstructure properties of B4C reinforced aluminium alloy composites. Mater. Sci. Technol. 2013, 29, 954–960. [Google Scholar] [CrossRef]
  9. Baradeswaran, A.; Perumal, A.E. Perumal, Influence of B4C on the tribological and mechanical properties of Al 7075-B4C composites. Compos. Part B 2013, 54, 146–152. [Google Scholar] [CrossRef]
  10. Lashgari, H.; Zangeneh, S.; Shahmir, H.; Saghafi, M.; Emamy, M. Heat treatment effect on the microstructure, tensile properties and dry sliding wear behavior of A356-10%B4C cast composites. Mater. Des. 2010, 31, 4414–4422. [Google Scholar] [CrossRef]
  11. Gudipudi, S.; Nagamuthu, S.; Subbian, K.S.; Prakasa, S.; Chilakalapalli, R. Enhanced mechanical properties of AA6061-B4C composites developed by a novel ultra-sonic assisted stir casting. Eng. Sci. Technol. Int. J. 2020, 23, 1233–1243. [Google Scholar] [CrossRef]
  12. Patidar, D.; Rana, R. Effect of B4C particle reinforcement on the various properties of aluminium matrix composites: A survey paper. Mater. Today Proc. 2017, 4, 2981–2988. [Google Scholar] [CrossRef]
  13. Moktar, M.S.; Ghandvar, H.; Abu Bakar, T. Microstructural and tensile properties of Al-20%Mg2Si-xSiCp hybrid metal matrix composite. Encycl. Renew. Sustain. Mater. 2020, 5, 54–63. [Google Scholar] [CrossRef]
  14. Sukiman, N.A.; Ghandvar, H.; Abu Bakar, T.; Wee, Y.C. Microstructure characterization and tensile properties of Al-15%Mg2Si-xYSZ hybrid composite. Malays. J. Microsc. 2019, 15, 30–37. [Google Scholar]
  15. Uvaraja, V.C.; Natarajan, N. Tribological characterization of stir-cast hybrid composite Aluminium 6061 reinforced with SiC and B4C particulates. Eur. J. Sci. Res. 2012, 76, 539–552. [Google Scholar]
  16. Aziz, M.A.; Mahmoud, T.S.; Zaki, Z.I.; Gaafer, A.M. Heat treatment and wear characteristics of Al2O3 and tic particulate reinforced AA6063 alloy hybrid composites. J. Tribol. 2006, 128, 891–904. [Google Scholar] [CrossRef]
  17. Benal, M.M.; Shivanand, H.K. Effects of reinforcement content and ageing durations on wear characteristics of Al6061 based hybrid composites. Wear 2007, 262, 759–763. [Google Scholar] [CrossRef]
  18. Umanath, K.; Selvamani, S.T.; Palanikumar, K. Friction and wear behaviour of Al6061 alloy (SiCp+ Al2O3) hybrid composites. Int. J. Eng. Sci. Technol. 2011, 3, 5441–5451. [Google Scholar]
  19. Altinkok, N.; Coban, A. The tensile behaviour and microstructure of Al2O3/SiCp reinforced aluminium-based mmcs produced by the stir casting method. Int. J. Adv. Sci. Technol. 2012, 2, 78–86. [Google Scholar]
  20. Uvaraja, V.C.; Natarajan, N.; Rajendran, I.; Sivakumar, K. Tribological behaviour of novel hybrid composite materials using Taguchi technique. J. Tribol. 2013, 135, 021101. [Google Scholar] [CrossRef]
  21. Sharma, S.S.; Jagannath, K.; Prabhu, P.R.; Gowri, S.; Harisha, S.R.; Kini, U.A. Metallography & bulk hardness of artificially aged Al6061-B4C-sic stir cast hybrid composites. Mater. Sci. Forum. 2017, 880, 140–143. [Google Scholar] [CrossRef]
  22. Uthayakumar, M.; Aravindan, S.; Rajkumar, K. Wear performance of Al-SiC-B4C hybrid composites under dry sliding conditions. Mater. Design. 2013, 47, 456–464. [Google Scholar] [CrossRef]
  23. Lin, Q.; Shen, P.; Qiu, F.; Zhang, D.; Jiang, Q. Wetting of polycrystalline B4C by molten Al at 1173–1473K. Scr. Mater. 2009, 60, 960–963. [Google Scholar] [CrossRef]
  24. Wu, H.; Zeng, F.; Yuan, T.; Zhang, F.; Xiong, X. Wettability of 2519Al on B4C at 1000–1250 °C and mechanical properties of infiltrated B4C-2519Al composites. Ceram. Int. 2014, 40, 2073–2081. [Google Scholar] [CrossRef]
  25. Shirvanimoghaddam, K.; Khayyam, H.; Abdizadeh, H.; Karbalaei, A.M.; Pakseresht, A.H.; Ghasali, E.; Naebe, M. Boron carbide reinforced aluminium matrix composite: Physical, mechanical characterization and mathematical modelling. Mater. Sci. Eng. A 2016, 658, 135–149. [Google Scholar] [CrossRef]
  26. Harichandran, R.; Selvakumar, N. Effect of nano/micro B4C particles on the mechanical properties of aluminium metal matrix composites fabricated by ultrasonic cavitation-assisted solidification process. Arch. Civ. Mech. Eng. 2015, 16, 147–158. [Google Scholar] [CrossRef]
  27. Mehmet, I.; Amin, N.; Onder, A.; Sabri, A. Mechanical characterization of B4C reinforced aluminum matrix composites produced by squeeze casting. Mater. Res. Soc. 2016, 32, 1–7. [Google Scholar] [CrossRef]
  28. Hashim, J.; Looney, L.; Hashmi, M.S.J. Metal matrix composites: Production by the stir casting method. J. Mater. Process. Technol 1999, 92, 1–7. [Google Scholar] [CrossRef]
  29. Bramfitt, B.L. The effect of carbide and nitride additions on the heterogeneous nucleation behavior of liquid iron. Metall. Trans. 1970, 1, 197–201. [Google Scholar] [CrossRef]
  30. Shabestari, S.G.; Malekan, M. Assessment of the effect of grain refinement on the solidification characteristics of 319 aluminum alloy using thermal analysis. J. Alloys. Compd. 2010, 492, 134–142. [Google Scholar] [CrossRef]
  31. Ghandvar, H.; Idris, M.H.; Ahmad, N.; Emamy, M. Effect of gadolinium addition on microstructural evolution and solidification characteristics of Al-15%Mg2Si in-situ composite. Mater. Char. 2018, 135, 57–70. [Google Scholar] [CrossRef]
  32. Nagarajan, S.; Dutta, B.; Surrapa, M.K. The effect of SiC particles on the size and morphology of eutectic silicon in cast A356/SiCp composites. Compos. Sci. Technol. 1999, 59, 897–902. [Google Scholar] [CrossRef]
  33. Li, Y.Z.; Wang, Q.Z.; Wang, W.G.; Xiao, B.L.; Ma, Z.Y. Interfacial reaction mechanism between matrix and reinforcement in B4C/6061Al composites. Mater. Chem. Phys. 2015, 1–11. [Google Scholar] [CrossRef]
  34. Azarbarmas, M.; Emamy, M.; Karamouz, M.; Alipour, M.; Rassizadehghani, J. The effects of boron additions on the microstructure, hardness and tensile properties of in situ Al-15%Mg2Si composite. Mater. Design. 2011, 32, 5049–5054. [Google Scholar] [CrossRef]
  35. Viala, J.C.; Bouix, J.; Gonzalez, G.; Esnouf, C. Chemical reactivity of aluminium with boron carbide. J. Mater. Sci. 1997, 32, 4559–4573. [Google Scholar] [CrossRef]
  36. Lashgari, H.R.; Emamy, M.; Razaghian, A.; Najimi, A.A. The effect of strontium on the microstructure, porosity and tensile properties of A356-10%B4C cast composite. Mater. Sci. Eng. A 2009, 517, 170–179. [Google Scholar] [CrossRef]
  37. Ghandvar, H.; Farahany, S.; Idris, J. Wettability Enhancement of SiCp in Cast A356/SiCp Composite Using Semisolid Process. Mater. Manuf. Process. 2015, 30, 1442–1449. [Google Scholar] [CrossRef]
  38. Song, M. Effects of volume fraction of SiC particles on mechanical properties of SiC/Al composites. Trans. Nonferrous Metals Soc. China 2009, 19, 1400–1404. [Google Scholar] [CrossRef]
  39. Qi, G.; Shusen, W.; Shulin, L.; Xuecheng, D.; Zhiyuan, Z. Preparation of in- situ TiB2 and Mg2Si hybrid particulates reinforced Al-matrix composites. J. Alloy. Compd. 2015, 651, 521–527. [Google Scholar] [CrossRef]
  40. Moses, J.; Dinaharan, I.; Sekhar, S. Prediction of influence of process parameters on tensile strength of AA6061/TiC aluminum matrix composites produced using stir casting. Trans. Nonferrous Met. Soc. China 2015, 26, 1498–1511. [Google Scholar] [CrossRef]
  41. Babout, L.; Maire, E.; Buffière, J.Y.; Fougères, R. Characterization by X-ray computed tomography of decohesion, porosity growth and coalescence in model metal matrix composites. Acta Mater. 2001, 49, 2055–2063. [Google Scholar] [CrossRef]
  42. Mohammadi, H.; Emamy, M.; Hamnabard, Z. The statistical analysis of tensile and compression properties of the as-cast AZ91-X%B4C composites. Inter. Metal. Cast. 2020, 14, 505–517. [Google Scholar] [CrossRef]
  43. Lakshmanan, P.A. Fabrication and Tribological Behaviour of (Mg-TiO2) Composites. Trans. Famena 2017, 41, 61–70. [Google Scholar] [CrossRef] [Green Version]
  44. Archard, J. Contact and rubbing of flat surfaces. J. Appl. Phys. 1953, 24, 981–988. [Google Scholar] [CrossRef]
  45. Thakur, S.K.; Dhindaw, B.K. The influence of interfacial characteristics between SiCp and Mg/Al metal matrix on wear, coefficient of friction and microhardness. Wear 2001, 247, 191. [Google Scholar] [CrossRef]
  46. Basavarajappa, S.; Chandramohan, G.; Mukund, K.; Ashwin, M.; Prabu, M. Dry sliding wear behavior of Al 2219/SiCp-Gr hybrid metal matrix composites. J. Mater. Eng. Perform. 2006, 15, 668. [Google Scholar] [CrossRef]
  47. Devaraju, A.; Pazhanivel, K. Evaluation of Microstructure, Mechanical and Wear Properties of Aluminium Reinforced with Boron Carbide Nano Composite. Indian. J. Sci. Technol. 2016, 9, 1–6. [Google Scholar] [CrossRef] [Green Version]
  48. Tang, F.; Wu, X.; Ge, S.; Ye, J.; Zhu, H.; Hagiwara, M.; Schoenung, J. Dry sliding friction and wear properties of B4C particulate-reinforced Al-5083 matrix composites. Wear 2008, 264, 555–561. [Google Scholar] [CrossRef]
Figure 1. (a) SEM image of Al-20%Mg2Si composite with B4C addition with (b) corresponding elemental mapping analysis.
Figure 1. (a) SEM image of Al-20%Mg2Si composite with B4C addition with (b) corresponding elemental mapping analysis.
Applsci 11 03047 g001
Figure 2. XRD results of Al-20%Mg2Si in situ composite added with 10 wt. % B4C.
Figure 2. XRD results of Al-20%Mg2Si in situ composite added with 10 wt. % B4C.
Applsci 11 03047 g002
Figure 3. SEM images of Al-20%Mg2Si composite with various additions of B4C: (a) 0 wt%, (b) 2.5 wt%, (c) 5 wt%, and (d) 10 wt%.
Figure 3. SEM images of Al-20%Mg2Si composite with various additions of B4C: (a) 0 wt%, (b) 2.5 wt%, (c) 5 wt%, and (d) 10 wt%.
Applsci 11 03047 g003
Figure 4. Characteristics of primary Mg2Si particles in Al-20%Mg2Si composite treated with various content of B4C.
Figure 4. Characteristics of primary Mg2Si particles in Al-20%Mg2Si composite treated with various content of B4C.
Applsci 11 03047 g004
Figure 5. Cooling curves with first and second derivative curves of the Al-20%Mg2Si composite added with various B4C concentrations: (a) 0, (b) 2.5, (c) 5, and (d)10 wt%.
Figure 5. Cooling curves with first and second derivative curves of the Al-20%Mg2Si composite added with various B4C concentrations: (a) 0, (b) 2.5, (c) 5, and (d)10 wt%.
Applsci 11 03047 g005
Figure 6. Nucleation temperature (TN) and growth temperature (TG) of primary Mg2Si phase correspond to different B4C additions.
Figure 6. Nucleation temperature (TN) and growth temperature (TG) of primary Mg2Si phase correspond to different B4C additions.
Applsci 11 03047 g006
Figure 7. Duration of primary Mg2Si particles to nucleate and grow as a result of B4C additions.
Figure 7. Duration of primary Mg2Si particles to nucleate and grow as a result of B4C additions.
Applsci 11 03047 g007
Figure 8. (a) SEM micrograph of primary Mg2Si crystal including B4C particle, (b) corresponding elemental mapping analysis.
Figure 8. (a) SEM micrograph of primary Mg2Si crystal including B4C particle, (b) corresponding elemental mapping analysis.
Applsci 11 03047 g008
Figure 9. Comparison of crystal structures of Mg2Si and B4C crystals.
Figure 9. Comparison of crystal structures of Mg2Si and B4C crystals.
Applsci 11 03047 g009
Figure 10. (a) BSE micrograph of Al-20%Mg2Si-10%B4C hybrid composite, (b) SEM image of Al-Mg-B compound, and (c) corresponding EDS result of the white particle in (b).
Figure 10. (a) BSE micrograph of Al-20%Mg2Si-10%B4C hybrid composite, (b) SEM image of Al-Mg-B compound, and (c) corresponding EDS result of the white particle in (b).
Applsci 11 03047 g010
Figure 11. Vickers hardness of Al-20%Mg2Si composite treated with various amount of B4C particles.
Figure 11. Vickers hardness of Al-20%Mg2Si composite treated with various amount of B4C particles.
Applsci 11 03047 g011
Figure 12. (a) Engineering stress-strain curve of Al-20%Mg2Si with various content of B4C (b) the ultimate tensile strength (UTS) and (c) percentage elongation (El%) values of Al-Mg2Si composite with different B4C contents.
Figure 12. (a) Engineering stress-strain curve of Al-20%Mg2Si with various content of B4C (b) the ultimate tensile strength (UTS) and (c) percentage elongation (El%) values of Al-Mg2Si composite with different B4C contents.
Applsci 11 03047 g012
Figure 13. Fractography of Al-20%Mg2Si tensile samples with different contents of B4C in low (left) and high (right) magnifications: (a,b) 0, (c,d) 5, and (e,f) 10 wt% B4C.
Figure 13. Fractography of Al-20%Mg2Si tensile samples with different contents of B4C in low (left) and high (right) magnifications: (a,b) 0, (c,d) 5, and (e,f) 10 wt% B4C.
Applsci 11 03047 g013
Figure 14. Wear rate of Al-Mg2Si composite reinforced with different amount of B4C particles under applied loads of 20 and 40 N.
Figure 14. Wear rate of Al-Mg2Si composite reinforced with different amount of B4C particles under applied loads of 20 and 40 N.
Applsci 11 03047 g014
Figure 15. Coefficient of friction of Al-20%Mg2Si composite reinforced with different amount of B4C particles.
Figure 15. Coefficient of friction of Al-20%Mg2Si composite reinforced with different amount of B4C particles.
Applsci 11 03047 g015
Figure 16. SEM micrographs of the worn surface of Al-Mg2Si composite samples with various B4C contents (a,b) 0 wt%, (c,d) 5 wt%, and (e,f) 10 wt% under 20 N (left) and 40 N (right) applied loads.
Figure 16. SEM micrographs of the worn surface of Al-Mg2Si composite samples with various B4C contents (a,b) 0 wt%, (c,d) 5 wt%, and (e,f) 10 wt% under 20 N (left) and 40 N (right) applied loads.
Applsci 11 03047 g016
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ghandvar, H.; Jabbar, M.A.; Koloor, S.S.R.; Petrů, M.; Bahador, A.; Bakar, T.A.A.; Kondoh, K. Role B4C Addition on Microstructure, Mechanical, and Wear Characteristics of Al-20%Mg2Si Hybrid Metal Matrix Composite. Appl. Sci. 2021, 11, 3047. https://doi.org/10.3390/app11073047

AMA Style

Ghandvar H, Jabbar MA, Koloor SSR, Petrů M, Bahador A, Bakar TAA, Kondoh K. Role B4C Addition on Microstructure, Mechanical, and Wear Characteristics of Al-20%Mg2Si Hybrid Metal Matrix Composite. Applied Sciences. 2021; 11(7):3047. https://doi.org/10.3390/app11073047

Chicago/Turabian Style

Ghandvar, Hamidreza, Mostafa Abbas Jabbar, Seyed Saeid Rahimian Koloor, Michal Petrů, Abdollah Bahador, Tuty Asma Abu Bakar, and Katsuyoshi Kondoh. 2021. "Role B4C Addition on Microstructure, Mechanical, and Wear Characteristics of Al-20%Mg2Si Hybrid Metal Matrix Composite" Applied Sciences 11, no. 7: 3047. https://doi.org/10.3390/app11073047

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