Next Article in Journal
Dynamic Simulation and Characteristic Analysis on Freezing Process in Ballast Tanks of Polar LNG Carriers
Previous Article in Journal
Research on Anti-Single-Frequency Interference Ability of Underwater Acoustic DSSS Communication System with Short-Period PN Sequences
Previous Article in Special Issue
Fracture Prediction in Weldox 700E Steel Subjected to High Velocity Impact Using LS-DYNA
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

In Situ Al3BC/Al Composite Fabricated via Solid-Solid Reaction: An Investigation on Microstructure and Mechanical Behavior

by
Tapabrata Maity
1,*,
Aditya Prakash
2,
Debdas Roy
1 and
Konda Gokuldoss Prashanth
3,4,*
1
Department of Materials and Metallurgical Engineering, National Institute of Advanced Manufacturing Technology (Previously NIFFT), Ranchi 834003, India
2
R & D Division, Tata Steel Ltd., Jamshedpur 831001, India
3
Department of Mechanical and Industrial Engineering, Tallinn University of Technology, Ehitajate Tee 5, 19086 Tallinn, Estonia
4
Center for Biomaterials, Cellular and Molecular Theranostics (CBCMT), School of Engineering, Vellore Institute of Technology, Vellore 632014, India
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2025, 15(9), 5189; https://doi.org/10.3390/app15095189
Submission received: 14 April 2025 / Revised: 4 May 2025 / Accepted: 5 May 2025 / Published: 7 May 2025

Abstract

:
Al3BC, with its remarkably high modulus of elasticity (326 GPa) and hardness (14 GPa), coupled with a low density (2.83 g/cc), stands out as a promising reinforcement material for Al matrix composite. To study feasibility of solid-solid reaction (SSR) by forming an in situ Al3BC reinforcing phase within the matrix, this study developed an Al3BC/Al composite via mechanical alloying, followed by sintering at 1000 °C/1 h, and subsequent hot pressing at 400 °C/40 MPa. The reaction kinetics and corresponding electron microscopy images suggest that the aluminum (Al)-boron (B) reacts with graphene nanoplates (GNPs) to form both clusters and a heterogeneous multi-structured Al3BC reinforcements network dispersed within the fine-grain (FG) Al matrix. The heterostructure contributes to a good balance between strength (~284 MPa) and ductility (~17%) and stiffness (~212 GPa). Superior strain hardening ability (n = 0.3515) endorses remarkable load-bearing capacity (σc = 1.63) and thereby promotes excellent strength-ductility synergy in the composite. The fracture morphology reveals that reasonable ductility primarily relies on the crack deflection by the FG-Al matrix, playing a critical role in delaying fracture. The potential importance of the matrix microstructure in the overall fracture resistance of the composite has been highlighted.

1. Introduction

Al-matrix composites (AMCs) offer lightweight solutions for the transport field, enabling significant progress in energy conservation and emission reduction. In addition, they boast a superior strength-to-weight ratio, a key property deeply desired in aerospace and military applications [1,2,3]. However, the challenge lies in balancing strength with ductility, despite the lightweight and durable nature of the Al matrix, enabling a serious threat to their adaptation in those applications [4]. While strength may increase with decreasing the grain size in the matrix in many nanostructured materials, at the same time it costs reduced ductility [5]. The limited ductility associated with nanostructured grain structures leads to a reduction in composite strain hardening capacity. On this occasion, Al-matrix composites with particulate reinforcement can typically provide several key advantages of superior specific strength and stiffness [6], enhanced wear resistance [7], and good structural integrity with property design flexibility [8]. Ceramic reinforcements with high modulus and strength, such as silicon carbide (SiC) [9], alumina (Al2O3) [10], boron carbide (B4C) [11], titanium boride (TiB2) [12], aluminum nitride (AlN) [13,14], zirconium boride (ZrB2) [15], titanium aluminide (Al3Ti) [16], magnesium silicide (Mg2Si) [17], etc., have been beneficial in enhancing the room temperature properties and softening resistance of these composites. According to research Al3BC (an Al compound)-reinforced AMCs (Al3BC/Al composites) show impressive properties: high modulus (100 GPa), strength (400 MPa), and reasonable ductility (6%) [18,19]. In addition, excellent Al3BC-Al matrix interfacial bonding promotes the development of a heterogeneous grain structure and boosts strengthening in the Al3BC/Al composites [20,21]. From a mechanical point of view, Al3BC has emerged as one of the potential candidates for reinforcing Al-matrix composites.
Al3BC is a promising reinforcement for the Al-matrix composites due to several advantages. Firstly, reinforcing Al3BC is significantly easier than conventional reinforcements. Additionally, it requires only minimal additions of B and C to obtain a high fraction of reinforcing Al3BC in the Al-B-C systems, even among other aluminum–boron carbide phases, e.g., Al3BC3, Al8B4C7, AlB24C4, etc. [22]. Secondly, Al3BC particles act as good heterogeneous nucleating sites for the matrix, as demonstrated by Ma et al. in Mg-ZA263 composites [22]. Notably, the unique distinctive crystal structure of Al3BC, rich in Al, offers an excellent combination of high hardness, like ceramic reinforcements, yet with a more metallic behavior [21]. Furthermore, the combination of properties of the Al3BC phase, such as low density (2.83 g/cc) compared to other reinforcements, e.g., Al2O3 (3.98 g/cc) and TiB2 (4.50 g/cc), excellent thermal stability (up to 1700 K under 1.6–4.8 GPa pressure), high hardness (14 GPa), and remarkably high modulus (Young’s modulus 326 GPa and shear modulus 140 GPa), makes it a promising candidate for reinforcing Al-matrix composites [21,22]. Zhao et al. demonstrated that Al-matrix composites reinforced with Al3BC particles show noticeably high room-temperature strength (775 MPa) and elastic modulus. Additionally, Al/Al3BC/25p composites, prepared via liquid–solid reaction (stir-casting) possess superior strain hardening ability [21]. Growing with such promise, Tian et al. and Zhao et al. explored remarkable strengthening in uniformly dispersed nano-Al3BC-reinforced 6061 Al-alloy and AZ63 Mg-alloys [23,24]. Due to their strength, machinability, and sustainable manufacturing demands, Al3BC/Al composites show promise for future-generation aerospace and automotive parts (airframes, propeller blades) [25].
Reinforcing graphene platelets (GNPs) uniformly in the Al matrix has been a decades-long challenge, which limits the development of graphene-reinforced high-strength Al-matrix composites. Recently, research by Mei et al. successfully demonstrated the fabrication of Al3BC/Al composites using a liquid reaction (stir casting) from amorphous B and GNPs [20]. This approach can potentially overcome the hurdle of graphene dispersion and pave the way a new generation of high-performance Al-matrix composites [26]. Powder metallurgy (P/M) offers a solid-solid reaction to create a bimodal heterostructure in the Al-matrix composites, using ball milling and high-press sintering to achieve homogeneous dispersion of reinforcements in the matrix [27,28,29,30,31,32,33]. The key advantage of solid–solid (pressing and sintering) reactions is the ability to control the morphology, shape, and size of the reinforcing phases that evolve during both in situ and ex situ approaches [17,34,35,36]. However, the solid-solid reaction path to develop in situ Al3BC/Al composites remain a challenge. Kubota et al. showed difficulty controlling microstructures with Al-MgB2 powder reactions because multiple by-products, like AlB2 and Al4C3, often resulted [37]. The presence of intermediate products, like micron-sized Al4C3, can significantly deteriorate the mechanical properties of Al3BC/Al composites [38]. Therefore, fabrication temperature and modulating the reaction products of Al3BC/Al composites are crucial to obtain the desired mechanical properties in these composites because of the effect of undesired reaction products. Accordingly, in this work, we focused on a straightforward method for fabricating Al3BC/Al composites using mechanical alloying (ball milling) combined with a solid-state reaction technique (pressing and sintering) by varying reaction time (3 to 7 h) at 1000 °C temperature to avoid unwanted byproducts with optimized mechanical properties.

2. Experimental Details

2.1. Materials

The investigated Al3BC/Al composite was fabricated via ball milling and vacuum sintering (1000 °C/1 h) to initiate solid–solid reaction of constituent elements of commercial Al powder (particle size 250–450 μm, HiMedia TM, Kennett Square, PA, USA), GNPs (average diameter 1.65 nm, Alfa Aesar TM, Haverhill, MA, USA) aggregates, and amorphous B (Sigma Aldrich TM, St. Louis, MO, USA). B and GNP powders (2:1 molar ratio) were mechanically alloyed in a high-energy ball mill (RetschTM PM 400-MA, Haan, Germany).

2.2. Processing

Milling was carried out for 2 h at a speed of 200 rpm in an Argon atmosphere to ensure uniform elemental distribution. The ball-milling parameters have been reported elsewhere [27]. The milled B+GNPs powder was mixed with Al powder (overall molar ratio 1:2:4) for an additional 5 h using a ball mill and subsequently pre-heated (overnight at 80 °C) and green compacted (60 vol.%). To prepare the corresponding Al3BC/Al composite, the green compact was then sintered in a vacuum sintering furnace (tubular furnace) at 1000 °C. The sintering process maintained a slower heating and pressurization process, e.g., heated at a rate of 5 °C/min, kept for 1 h (soaking time) below 5 × 10−2 mbar pressure, and then slowly cooled to room temperature. This ensures Al will maintain formability favorably and even become liquid at 680 °C. Finally, a low-pressure (40 MPa) and high-temperature (400 °C) hot-press deformation was applied for 20 min in a graphite mold (φ 20 mm) to achieve the final consolidated specimens. During hot pressing, the heating rate was maintained at 25 °C/min, followed by water cooling. The experimental procedure employed in this research is shown schematically in Figure 1.

2.3. Characterization

The composite samples were subjected to standard metallurgical procedures, including grinding and polishing, and subsequently cut (as cylindrical compression test specimens) using wire EDM following standard procedure. The overall phase composition was evaluated with X-ray diffraction (XRD) using RigakuTM DMAX X-ray diffractometer [operating at 40 kV-40 mA and Cu-Kα radiation] (Tokyo, Japan). The reactions during the composite formation were investigated using differential scanning calorimetry (DSC). Microstructure and composition area mapping were then investigated using an electron microscope (SEM) equipped with energy-dispersive X-ray spectroscopy (EDS). The SEM device fitted with an electron backscattered diffraction (EBSD) detector was used to obtain inverse pole figure maps (IPF) with step size 0.5 μm by an EDAX TridentTM system (Sparks Glencoe, MD, USA) along with system-integrated FEI Scios™ field emission gun (FEG) scanning electron microscope (FE-SEM) (Sunnyvale, CA, USA). The EBSD data were processed using an EDAX OIM Analysis™ software (https://www.edax.com.cn/products/ebsd/oim-analysis, accessed on 4 May 2025) and a detailed description of the technique has been reported elsewhere [27]. The hardness was measured using a Vickers hardness tester at a load of 100 g [ASTM E10-14 standard [39]]. The room temperature (RT) compression test (strain rate ( ε ˙ ) of 0.5 × 10−3 s−1) specimens with dimensions 20 × 5.6 × 5 mm3 were evaluated using a servo-hydraulic UTM (100 kN load cell) [ASTM-E08 standard [40]]. The tests continued until the failure of the specimen. The fracture surfaces of the post-compressed specimens were examined using SEM to understand the mode of failure.

3. Results

3.1. Physical and Structrual Characterization

Figure 2a presents the DSC traces of the ball-milled Al and B+GNP powders conducted at a heating rate of 5 °C/min under an argon atmosphere. An exothermic peak is observed beyond the melting point of Al (660 °C) around 680 °C [41], signifying the possible chemical reaction between Al, B, and GNPs initiating solid-solid reactions [21]. No (endothermic/exothermic) peaks are observed beyond 1000 °C. This suggests that there is a simultaneous reaction of Al, B, and GNPs and the in-situ synthesis of Al3BC. The XRD patterns (Figure 2b) of (B+GNPs)/Al milled powders reveal the presence of crystalline fcc-Al phase [space group: Fm-3m; PDF-2-2023#00-004-0787]. The peak broadening with increasing milling time indicates the possible structural refinement. Furthermore, a decrease in intensity of the fcc-Al peak (111) (w.r.t. (200) peak) demonstrates the dissolution of the fcc phase, which indicates successful alloying [42]. Figure 2c shows the XRD pattern of the as-fabricated Al3BC/Al composite. The diffraction peaks show the presence of fcc-Al, hcp-Al3BC [space group: P63/mmc; PDF-2-2023#00-047-1628], and hcp-AlB2 [space group: P6/mmm; PDF-2-2023#03-65-3381] phases. Importantly, no unwanted Al4C3, a common by-product during solid-liquid or liquid-liquid state reaction(s), was detected. In addition, a higher intensity of Al and Al3BC phases than AlB2 phases indicates their dominant behavior. The lattice parameters (a), and the relative amounts of the phases in terms of mass fraction/phase volume fractions, Vf, were determined using the XRD peak profile analysis (refer to Table 1 for details) [42]. It shows the volume fraction of the Al matrix ( V f A l ) accounts for 25 ± 2 wt.%, primary reinforcement Al3BC ( V f A l 3 B C ) corresponds to 65 ± 2 wt.%, and AlB2 ( V f A l B 2 ) volume fraction is calculated to be 10 ± 1 wt.%. This suggests that the addition of GNPs and B promotes the Al3BC formation preferentially reacting with Al over forming AlB2.

3.2. Microstructure Evolution and Reinforcement Distribution

Figure 3a shows a typical composite microstructure of the as-fabricated Al3BC/Al specimen with heterogeneous grain structure comprised of both coarse (CGs) and fine grains (FGs). The microstructure reveals a deformed morphology with the presence of the following phases: a denser and wider size distribution of elongated shaped (60–70 µm) agglomerates (clusters) of grey Al3BC, darker contrast AlB2 phases, and a string/or network of submicron grey Al3BC phases dispersed within a highly deformed (brighter contrast) fcc-Al matrix (Figure 3b). The EDS element overlay mapping in Figure 3c identifies the following elements: B, C, and Al. The EDS results confirm the presence of the desired in situ Al3BC phase (yellow circle), and the matrix is primarily composed of fcc-Al. Notably, network-like Al3BC structure comprises both CG (elongated) and FG (equiaxed) grains and clusters of black, rod-like, micron-size AlB2 phases located at the interfaces of Al3BC agglomerates. However, obtaining a clear element overlay for the AlB2 phase was challenging due to its brittle nature.
The standard metallographic preparation employed likely fractured the clusters, leading to the presence of inclusions. Similar morphology can be found in other relevant studies [38]. Therefore, the overall microstructure reveals a heterogeneous microstructure where in situ Al3BC phase with varying grain sizes, from coarse-grained Al3BC to sub-micron fine-grained Al3BC, dispersed within fcc-Al matrix [43]. The solid-solid reaction at 1000 °C likely accelerated the agglomeration, promoting the formation of micron-scale Al3BC/AlB2 clusters within the composite [44]. The phase area fraction (Af) was quantitatively measured (including more than 5000 particles) using the ImageJ software (https://imagej.net/ij/). The average Af values were found to be A f A l 2 O 3 = 52 ± 4%, A f A l = 31 ± 2%, and A f A l B 2 = 17 ± 3%, respectively. Such heterogeneity in microstructure is in good agreement with the XRD results [ V f A l 3 B C = ~65 wt.%, V f A l = ~25 wt.%, and V f A l B 2 = ~10 wt.%] (Table 1). The presence of micro-voids (Figure 3b), cracks (Figure 3a), and fracturing in some CG-Al3BC phases can be observed throughout the microstructure, possibly due to the low pressure and high temperature (400 °C/40 MPa) employed during deformation (hot press).

3.3. Grain Structure Evolution

Figure 4 shows the phase fraction, orientation, and inverse pole figure maps (IPF) obtained from the EBSD analysis of the corresponding SEM micrograph shown in Figure 4a. Figure 4b reveals that the matrix consists of near equiaxed FG-Al (green), FG-Al3BC (red), and elongated CG-Al3BC (red) phases that arrange themselves heterogeneously throughout the microstructure. Such microstructural formation is likely due to ongoing deformation, showing a heterogeneous structure. Here, grain size, smaller than 2 µm is defined as FG. FG-Al occupies most of the visual field area and becomes the matrix phase, leading to 41% of the overall composition being FG-Al, with around 19% FG/CG-Al3BC phases being present within the FG-Al matrix. Additionally, the darker regions, along with the red areas in the map, likely correspond to the unindexed AlB2 phase. Such heterogeneous microstructures indicates the possible dynamic recrystallization, which usually happens during thermal processing processes [45,46,47]. Comparing the orientation map with the phase map, the recrystallized fcc-Al (FG-Al) occupying the FG regions, and the CG regions occupied by the deformed Al3BC phase can be observed.
According to the IPF image, the orientation of FG-Al (green), FG-Al3BC (red), and elongated CG-Al3BC (red) phases is random and the misorientation between them is high-angle. Similar phenomena were observed in Al3BC/6061 Al composites, where a dynamic recrystallization-induced fine-grain structure was observed in the Al matrix, which is likely due to dislocation accumulation via plastic deformation by extrusion [18,19]. Therefore, the present results suggest that hot pressing significantly initiates dislocation accumulation and local plastic deformation (SGB deformation) in fcc-Al grains. Further deformation refines the Al grains into near-equiaxed FG-Al grains and brittle agglomerated Al3BC reinforcements into FG-Al3BC within the FG-Al matrix, thereby leading to a heterogeneous structure within the composite. Interestingly, during thermal processing, such heterogeneous hard FG-Al3BC and FG-Al arrangements are likely to limit/hinder migration of grain boundaries and promote mechanical incompatibility at regional boundaries between FG-FGs and FG-CGs, resulting in a strain gradient.

3.4. Evolution of Mechanical Properties

3.4.1. Compressive Response

The room temperature compression test was used to calculate the Young’s modulus (E), yield strength (σy), ultimate compressive stress (σs), compressive plastic strain (εs), and strain hardening capacity (σsy) of the composite (Table 2). The compression engineering stress-strain curve displayed in Figure 5a exhibits Young’s modulus (E) of 212 GPa, a uniform deformation with σy of 108 MPa and σs of 284 MPa without sacrificing plasticity (εs) of 17%. A close inspection shows a significant strain hardening capacity (σsy) of up to 176 MPa in the composite, enabling good load-bearing capacity during plastic deformation and delaying the failure. In addition, serrated plastic flow (Luders’ strain) can be observed in the strain-hardened regions (marked in Figure 5a). Interestingly, such observed capacity (σsy) is higher than the values observed for TiB2 (1–4 vol.%)-reinforced Ti matrix composites showing (σsy) = 105–113 MPa [48]. The work hardening capacity (σc) was measured as (σc = (σs/σy) − 1) 1.63, indicating better load-bearing ability in the composite than other particulate reinforced composites (Table 2) [49,50]. Furthermore, σc is strongly correlated to its yield strength. This relationship is governed by the Hall-Petch relationship, which considers both grain size and dislocation density contrast factor between adjacent grains [49]. Therefore, it appears that an increase in the grain size, i.e., hard CG/clusters reinforcements, can significantly lower the σy (easier to deform initially) but has a greater potential to increase the strain hardening capacity σc in Al3BC/Al composites [50]. A close-up of the engineering stress–strain diagram reveals a serrated plastic flow, which indicates potential key occurrences of the formation of high-density dislocations, and initiating dislocation plasticity events [23,24].

3.4.2. Pronounced Strain Hardening Behavior

Strain hardening behavior of the composite was analyzed and plotted in Figure 5b by measuring the strain hardening exponent (n) using Hollomon equations:
σ = K ε n ,
n = d ( l n σ ) / d ( l n ε ) ,
where σ is stress, ε is strain, and K is a constant [52]. The plot shows strain hardening behavior depending on the microstructure of the composite. It reveals two stages of strain hardening, e.g., an initial high rate due to hard reinforcements (n1 = 0.4112) and a later more homogeneous deformation due to particulate matrix involvement interaction (n2 = 0.3048), with an overall strain hardening rate observed as 0.3515. It is reasonable that particulate matrix involvement interaction (n2 = 0.3048) influences the serrated plastic flow and promotes uniform deformation and superior ductility by delaying localized failure.
Serrated plastic flow appears either due to displacive phase transformation or the initiation of dislocation plasticity events [53,54]. As the XRD ruled out the possibilities of displacive phase transformation, dislocation plasticity events likely govern the mode of plastic deformation in Al3BC/Al composite. Furthermore, such serrated flow resembles Type-B serrations, which seemingly pronounce dislocation interaction involvement during SGB deformation [53,54]. Vickers hardness was measured to determine the individual strength of the phases and their influence on the overall strength of the composites (Table 2). It shows significantly higher hardness for Al3BC (CG) of HAl3BC = 22 ± 7 GPa, compared to FG-Al matrix, HAl = 3 ± 1 GPa, and AlB2, HAlB2 = 2 ± 0.5 GPa. The measured HAl3BC values are identical to the measured hardness values of Al3BC from other published reports, e.g., HAl3BC = 24 ± 1 [55], ~26 [21], respectively. Surprisingly, the hardness of the FG-Al matrix within the composite surpasses that of conventional Al (~1 GPa) [18,19,20]. Voigt’s model predicted the theoretical Young’s modulus (ETh) of the composite [56]:
E T h = f p E p + f m E m ,
where fp and fm are the volume fraction of the particles and the matrix and Ep and Em correspond to the Young’s modulus of Al3BC (326 GPa) [18,19] and FG-Al (52 GPa), as measured using nanoindentation, respectively. The measured Young’s modulus of FG-Al matches the literature value of ~56 GPa [21]. The fitted Young’s modulus ETh = 224 GPa matches the experimental Young’s modulus of the Al3BC/Al composite (E = 12 GPa). Furthermore, bulk modulus (G = 80 GPa) and shear modulus (B = 196 GPa) were calculated from the empirical equations [57].
The hardness (HTh) of the Al3BC/Al composite was theoretically predicted using the ROM [58]:
H T h _ u p = f p H p + f m H m ,
H T h _ l o w = f p H p + f m H m 1 ,
where Hp and Hm are the Vickers hardness, and fp and fm correspond to the mass fraction of Al3BC particle and Al matrix (obtained using the XRD patterns), respectively. Notably, the AlB2 phase was ignored due to low contributions in hardness (HAlB2 = 2 ± 0.5 GPa) and mass fraction (low V f A l B 2 = 10 ± 0.1 wt.%). Such significantly improved mechanical properties surpass those of deformed Al alloys (Table 2).

3.4.3. Strengthening Parameters Measurements

Apparently, strain hardening leads to increased hardness, modulus, and overall strength, while still maintaining good ductility. In this case, the achieved mechanical property may be compared with Al3BC/Al-Cu composites containing 13% and 26% Al3BC particles in the Al matrix, which possess high tensile strength (300 MPa–427 MPa) but lower elongations (2–4.5%) [53]. This justifies further investigation of strengthening parameters.
  • Load-bearing contribution (ΔσLoad): The strength increment due to load-bearing capacity of Al3BC reinforcement was determined from the expression [59]:
    σ L o a d = A + 1 4 υ P σ m ,
    where υp denotes the volume fraction of Al3BC particles (υp = 0.65) and σm refers to the yield strength of the composite and the particle aspect ratio denoted by A (~1) [18,19]. Hence, the contribution ΔσLoad of Al3BC reinforcement was predicted to be ΔσLoad = 12 MPa, comparable to ΔσLoad = 16 MPa in the AlSi10Mg/SiC (10 wt.%) matrix composite, exhibiting UTS 350 MPa and elongation to failure up to ~3% [60].
  • Strengthening capability contribution (R): The contribution of a specific volume fraction of Al3BC reinforcements to the strengthening of the Al matrix in Al3BC/Al composite, R, was evaluated from [61]:
    R = ( σ y c σ y m ) / f p σ y m
    The strengthening capability of the Al3BC/Al composites, calculated as R = 5.6 is impressive. The value is close to FeSiB-reinforced Cu matrix composites (R = 5.9) [62] and even surpasses SiC-particle-reinforced (R = 2.5) [63] and Al2O3-particle-reinforced (R = 2.3) [48] Cu matrix composites by a factor of two. This finding highlights the remarkable strengthening efficiency achieved in Al3BC/Al composites.
  • Furthermore, Pugh’s ratio, which is the ratio of the calculated bulk modulus (G = 80 GPa) to shear modulus (B = 196 GPa) (B/G), can be used to determine the ductility or brittleness of the material. A Pugh’s ratio greater than 1.75 indicates that the material is more ductile in nature [64]. Based on the calculated Pugh’s ratio, a B/G = 2.45 is observed for the as-developed Al3BC/Al composite. It can be concluded that the as-fabricated Al3BC/Al composite has superior ductility while considering the poison’s ratio of υ = 0.32. Furthermore, if the Vickers hardness of a material falls within the range of 10–20 GPa, it is typically referred to as hard [65]. According to this information, the composite (HTh = 15 ± 5 GPa) is categorized as a hard material.

4. Discussion

4.1. Fracture Surface Investigation

To better understand the strengthening effect of particles and the failure mechanism of the corresponding composite, the fracture surface of the sample after a compression test has been investigated. Figure 6a shows a typical compression fracture surface, which reveals primary cracks following the Al3BC network, with limited penetration in the matrix. AlB2 and Al3BC phases are exposed on the fractured surface because of their low ductility, and cracks appear at interfaces with the matrix. Figure 6b reveals that Al3BC particles have a hexagonal morphology with a sub-micron size. The broken particles highlighted by arrows further support their greater load-bearing ability, like mechanically milled and hot-extruded Mg2Si/Al composites [17]. In addition, Figure 6b reveals that the FG-Al phases are tightly surrounded by the Al matrix (white circle). Figure 6c reveals that the traces of dimples (red arrow) significantly suggest the ductile nature of fractures in FG-Al grains. Furthermore, Figure 6c shows the tear ridges (yellow arrow), which indicate localized plastic deformation around fractured particles, further supporting some ductility, along with cleavage planes (white arrow). These observations imply a mixed mode of fracture, which identifies both tough and brittle characteristics that exist in Al3BC/Al composites.

4.2. Understanding the Strength-Ductility Response in the Composite

The cracked Al3BC phase (Figure 6b) supports its brittleness (Figure 3c). Therefore, the fracture morphology suggests that the large CG/clusters might be the origin of the failure in these composites. Traces of pull-out (indicating interfacial debonding, suggesting Al3BC particles might have reached their strength limit) and tear ridges (indicating stress concentration occurring around these particles) (Figure 6b,c) suggest Al3BC bears significant load during deformation and stress would concentrate around these particles. Brittle FG/CG-Al3BC (Figure 6c) initiates cracks or deboning from the matrix, which ultimately leads to the formation of primary cracks (Figure 6a). In addition, Figure 6c reveals some micron-sized AlB2 particles fractured with smooth surfaces, while some particles remain intact but deboned from the matrix, supporting the brittleness claims. Figure 6d shows that while the primary crack propagates along interfaces, a few irregular secondary cracks initiate within the FG-Al matrix and penetrate further, suggesting some resistance to fracture. These secondary cracks (Figure 6d, yellow arrow) are ultimately bridged and deflected by the surrounding FG-Al matrix, delaying the fracture. This highlights the potential role of the FG-Al matrix in enhancing fracture resistance. As the Al3BC phase has a high-volume fraction in the composite, the inner unit stress facilitates load transfer to the adjacent units (n = 0.4112) to coordinate the deformation, likely contributing to early yielding of as-fabricated Al3BC/Al composites.
Continued loading promotes crack bridging through superplastic flow, maintaining a global strain hardening (n = 0.3048) and resulting in adequate ductility (17%). A high-volume fraction of brittle and hard Al3BC ( V f A l 3 B C = 65 ± 0.1 wt.%) transfers load to the adjacent unit to coordinate the deformation, ultimately improving strength and ductility. However, a very high amount of Al3BC ( V f A l 3 B C = 65 ± 0.1 wt.%) leads to an increase in the probability of primary crack formation and causes the early fracture in the compression test. Such a distinctive microstructure arrangement is beneficial to strengthening the composite, which is favorable to the strength and ductility of the composite.

5. Conclusions

In summary, the as-fabricated Al3BC/Al composites involving solid-state reactions exhibit improved mechanical properties. These reaction kinetics suggest that that prolonged milling duration (7 h) accelerates the solid–solid reaction to initiate in the modified alloy. Structure and microstructure investigations reveal that hot pressing (400 °C/40 MPa) effectively promotes a multi-scale heterostructure in the composite via size reduction by breaking down and dispersing agglomerates within the composite from coarse-grain CG-Al3BC to sub-micron FG-Al3BC dispersed within the FG-Al matrix. Strong interfaces between the FG-Al matrices promote crack bridging deflection through the plastic flow, reducing crack propagation, delaying fracture, and ultimately maintaining adequate ductility.

Author Contributions

Conceptualization, T.M. and K.G.P.; methodology, T.M., A.P. and D.R.; validation, T.M., D.R. and K.G.P.; formal analysis, T.M. and K.G.P.; investigation, T.M., A.P., D.R. and K.G.P.; resources, T.M., A.P. and D.R.; data curation, T.M. and K.G.P.; writing—original draft preparation, T.M., A.P. and D.R.; writing—review and editing, T.M. and K.G.P.; visualization, T.M. and K.G.P.; project administration, T.M. and K.G.P.; funding acquisition, T.M., D.R. and K.G.P. All authors have read and agreed to the published version of the manuscript.

Funding

The authors would like to thankfully acknowledge financial support provided by DST-SERB, New Delhi, project no. CRG/2020/005600.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

Author Aditya Prakash was employed by the company Tata Steel Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Das, D.K.; Mishra, P.C.; Singh, S.; Thakur, R.K. Properties of Ceramic-Reinforced Aluminium Matrix Composites—A Review. Int. J. Mech. Mater. Eng. 2014, 9, 12. [Google Scholar] [CrossRef]
  2. Tang, S.; Ummethala, R.; Suryanarayana, C.; Eckert, J.; Prashanth, K.G.; Wang, Z. Additive Manufacturing of Aluminum-Based Metal Matrix Composites—A Review. Adv. Eng. Mater. 2021, 23, 2100053. [Google Scholar] [CrossRef]
  3. Zhang, Z.; Wang, Z.; Zhao, Q.; Prashanth, K.G. Metal-Metal Interpenetrating Phase Composites: A Review. J. Alloys Compd. 2024, 1009, 176951. [Google Scholar] [CrossRef]
  4. Suryawanshi, J.; Prashanth, K.G.; Scudino, S.; Eckert, J.; Prakash, O.; Ramamurty, U. Simultaneous Enhancements of Strength and Toughness in an Al-12Si Alloy Synthesized Using Selective Laser Melting. Acta Mater. 2016, 115, 285–294. [Google Scholar] [CrossRef]
  5. Milligan, W.W.; Milne, I.; Ritchie, R.O.; Karihaloo, B. Comprehensive Structural Integrity; Milligan, W.W., Milne, I., Ritchie, R.O., Karihaloo, B., Eds.; Elsevier Science: Pergamon, Türkiye, 2003; Volume 1, ISBN 978-0-08-043749-1. [Google Scholar]
  6. Chawla, N.; Shen, Y.-L. Mechanical Behavior of Particle Reinforced Metal Matrix Composites. Adv. Eng. Mater. 2001, 3, 357–370. [Google Scholar] [CrossRef]
  7. Liddicoat, P.V.; Liao, X.Z.; Zhao, Y.; Zhu, Y.; Murashkin, M.Y.; Lavernia, E.J.; Valiev, R.Z.; Ringer, S.P. Nanostructural Hierarchy Increases the Strength of Aluminium Alloys. Nat. Commun. 2010, 1, 1–7. [Google Scholar] [CrossRef] [PubMed]
  8. Ibrahim, I.A.; Mohamed, F.A.; Lavernia, E.J. Particulate Reinforced Metal Matrix Composites—A Review. J. Mater. Sci. 1991, 26, 1137–1156. [Google Scholar] [CrossRef]
  9. Mayer, C.R.; Yang, L.W.; Singh, S.S.; Llorca, J.; Molina-Aldareguia, J.M.; Shen, Y.L.; Chawla, N. Anisotropy, Size, and Aspect Ratio Effects on Micropillar Compression of AlSiC Nanolaminate Composites. Acta Mater. 2016, 114, 25–32. [Google Scholar] [CrossRef]
  10. Slipenyuk, A.; Kuprin, V.; Milman, Y.; Goncharuk, V.; Eckert, J. Properties of P/M Processed Particle Reinforced Metal Matrix Composites Specified by Reinforcement Concentration and Matrix-to-Reinforcement Particle Size Ratio. Acta Mater. 2006, 54, 157–166. [Google Scholar] [CrossRef]
  11. Polat, S.; Sun, Y.; Çevik, E.; Colijn, H. Microstructure and Synergistic Reinforcing Activity of GNPs-B4C Dual-Micro and Nano Supplements in Al-Si Matrix Composites. J. Alloys Compd. 2019, 806, 1230–1241. [Google Scholar] [CrossRef]
  12. Xiao, P.; Gao, Y.; Yang, C.; Li, Y.; Huang, X.; Liu, Q.; Zhao, S.; Xu, F.; Gupta, M. Strengthening and Toughening Mechanisms of Mg Matrix Composites Reinforced with Specific Spatial Arrangement of In-Situ TiB2 Nanoparticles. Compos. B Eng. 2020, 198, 108174. [Google Scholar] [CrossRef]
  13. Sun, X.; Fan, Y.; Nie, J.; Chen, Y.; Xie, K.; Liu, S.; Zhao, Y.; Liu, X. Significant Improvement of the Room and Cryogenic Mechanical Properties of an AlN Particle Reinforced Al Matrix Composite by Alloying Element Magnesium. Compos. B Eng. 2024, 268, 111056. [Google Scholar] [CrossRef]
  14. Scudino, S.; Ali, F.; Surreddi, K.B.; Prashanth, K.G.; Sakaliyska, M.; Eckert, J. Al-Based Metal Matrix Composites Reinforced with Nanocrystalline Al-Ti-Ni Particles. J. Phys. Conf. Ser. 2010, 240, 012154. [Google Scholar] [CrossRef]
  15. Guan, C.; Chen, G.; Kai, X.; Cao, R.; Miao, C.; Xu, Z.; Zhao, Y. Evolution of Microstructure and Mechanical Properties of Graphene Nanoplates and ZrB2 Nanoparticles Reinforced AA6111 Composites during Hot Rolling Deformation. J. Alloys Compd. 2022, 920, 165910. [Google Scholar] [CrossRef]
  16. Geng, H.; Cui, C.; Liu, L.; Liang, Y. The Microstructures and Mechanical Properties of Hybrid In-Situ AlN-TiC-TiN-Al3Ti/Al Reinforced Al-Cu-Mn-Ti Alloy Matrix Composites. J. Alloys Compd. 2022, 903, 163902. [Google Scholar] [CrossRef]
  17. Yi, L.F.; Kunimoto, S.; Ishii, T.; He, L.; Onda, T.; Chen, Z.C. Improved Mechanical Properties of Mechanically Milled Mg2Si Particles Reinforced Aluminum-Matrix Composites Prepared by Hot Extrusion. Mater. Sci. Eng. A 2023, 871, 144904. [Google Scholar] [CrossRef]
  18. Zhao, Y.; Singaravelu, A.S.S.; Ma, X.; Zhang, Q.; Chang, S.L.Y.; Liu, X.; Chawla, N. Unveiling the Deformation Behavior and Strengthening Mechanisms of Al3BC/Al Composites via in-Situ Micropillar Compression. J. Alloys Compd. 2020, 823, 153842. [Google Scholar] [CrossRef]
  19. Zhao, Y.; Singaravelu, A.S.S.; Ma, X.; Zhang, Q.; Liu, X.; Chawla, N. Micromechanical Properties and Deformation Behavior of Al3BC/6061 Al Composites via Micropillar Compression. Mater. Sci. Eng. A 2020, 773, 138852. [Google Scholar] [CrossRef]
  20. Mei, Y.; Li, H.; Yang, W.; Wu, J.; Li, X.; Xiu, Z.; Fu, J.; Hussain, M.; Chen, G.; Wu, G. In-Situ Synthesis of Al3BC/Al Composites from Amorphous Boron and Graphene Nanoplates by Solid Reaction. J. Alloys Compd. 2020, 832, 154912. [Google Scholar] [CrossRef]
  21. Zhao, Y.; Singaravelu, A.S.S.; Ma, X.; Liu, X.; Chawla, N. Mechanical Properties of Al3BC by Nanoindentation and Micropillar Compression. Mater. Lett. 2020, 264, 127361. [Google Scholar] [CrossRef]
  22. Han, B.Q.; Huang, J.Y.; Zhu, Y.T.; Lavernia, E.J. Strain Rate Dependence of Properties of Cryomilled Bimodal 5083 Al Alloys. Acta Mater. 2006, 54, 3015–3024. [Google Scholar] [CrossRef]
  23. Tian, W.; Li, P.; Liu, X. Morphology Stability of Al3BC Phase in Aluminum Alloys. J. Alloys Compd. 2014, 583, 329–334. [Google Scholar] [CrossRef]
  24. Zhao, Y.; Ma, X.; Chen, H.; Zhao, X.; Liu, X. Preferred Orientation and Interfacial Structure in Extruded Nano-Al3BC/6061 Al. Mater. Des. 2017, 131, 23–31. [Google Scholar] [CrossRef]
  25. Lee, C. Do Effects of Microporosity on Tensile Properties of A356 Aluminum Alloy. Mater. Sci. Eng. A 2007, 464, 249–254. [Google Scholar] [CrossRef]
  26. Prashanth, K.G.; Scudino, S.; Chaubey, A.K.; Löber, L.; Wang, P.; Attar, H.; Schimansky, F.P.; Pyczak, F.; Eckert, J. Processing of Al-12Si-TNM Composites by Selective Laser Melting and Evaluation of Compressive and Wear Properties. J. Mater. Res. 2016, 31, 55–65. [Google Scholar] [CrossRef]
  27. Roy, D.; Pal, T.; Ajay, S.; Prakash, A.; Dutta, S.; Maity, T. Improving Strain Hardening Behavior in Nano-Intermetallic Reinforced Aluminum in-Situ Composites through an Optimized Twostep Thermal Processing Method; Sintering and Uniaxial Forging. J. Alloys Compd. 2024, 982, 173688. [Google Scholar] [CrossRef]
  28. Wang, Z.; Qu, R.T.; Scudino, S.; Sun, B.A.; Prashanth, K.G.; Louzguine-Luzgin, D.V.; Chen, M.W.; Zhang, Z.F.; Eckert, J. Hybrid Nanostructured Aluminum Alloy with Super-High Strength. NPG Asia Mater. 2015, 7, 1–8. [Google Scholar] [CrossRef]
  29. Markó, D.; Prashanth, K.G.; Scudino, S.; Wang, Z.; Ellendt, N.; Uhlenwinkel, V.; Eckert, J. Al-Based Metal Matrix Composites Reinforced with Fe49.9Co35.1Nb7.7B4.5Si2.8 Glassy Powder: Mechanical Behavior under Tensile Loading. J. Alloys Compd. 2015, 615, S382–S385. [Google Scholar] [CrossRef]
  30. Surreddi, K.B.; Scudino, S.; Sakaliyska, M.; Prashanth, K.G.; Sordelet, D.J.; Eckert, J. Crystallization Behavior and Consolidation of Gas-Atomized Al84Gd6Ni7Co3 Glassy Powder. J. Alloys Compd. 2010, 491, 137–142. [Google Scholar] [CrossRef]
  31. Prashanth, K.G.; Scudino, S. Quasicrystalline Composites by Additive Manufacturing. Key Eng. Mater. 2019, 818, 72–76. [Google Scholar] [CrossRef]
  32. Prashanth, K.G.; Kumar, S.; Scudino, S.; Murty, B.S.; Eckert, J. Fabrication and Response of Al70Y16Ni 10Co4 Glass Reinforced Metal Matrix Composites. Mater. Manuf. Process. 2011, 26, 1242–1247. [Google Scholar] [CrossRef]
  33. Xi, L.; Gu, D.; Guo, S.; Wang, R.; Ding, K.; Prashanth, K.G. Grain Refinement in Laser Manufactured Al-Based Composites with TiB2 Ceramic; Elsevier: Amsterdam, The Netherlands, 2020; Volume 9, pp. 2611–2622. [Google Scholar]
  34. Wang, Z.; Prashanth, K.G.; Scudino, S.; Chaubey, A.K.; Sordelet, D.J.; Zhang, W.W.; Li, Y.Y.; Eckert, J. Tensile Properties of Al Matrix Composites Reinforced with in Situ Devitrified Al84Gd6Ni7Co3 Glassy Particles. J. Alloys Compd. 2014, 586, 419–422. [Google Scholar] [CrossRef]
  35. Wang, Z.; Tan, J.; Scudino, S.; Sun, B.A.; Qu, R.T.; He, J.; Prashanth, K.G.; Zhang, W.W.; Li, Y.Y.; Eckert, J. Mechanical Behavior of Al-Based Matrix Composites Reinforced with Mg 58Cu28.5Gd11Ag2.5 Metallic Glasses. Adv. Powder Technol. 2014, 25, 635–639. [Google Scholar] [CrossRef]
  36. Wang, Z.; Prashanth, K.G.; Chaubey, A.K.; Löber, L.; Schimansky, F.P.; Pyczak, F.; Zhang, W.W.; Scudino, S.; Eckert, J. Tensile Properties of Al-12Si Matrix Composites Reinforced with Ti-Al-Based Particles. J. Alloys Compd. 2015, 630, 256–259. [Google Scholar] [CrossRef]
  37. Kubota, M.; Cizek, P. Synthesis of Al3BC from Mechanically Milled and Spark Plasma Sintered Al–MgB2 Composite Materials. J. Alloys Compd. 2008, 457, 209–215. [Google Scholar] [CrossRef]
  38. Li, D.; Hu, K.; Han, M.; Xie, K.; Liu, G.; Liu, X. Strength Adjust Mechanism of High Toughness Al4C3/Al-XMg Composites. Mater. Sci. Eng. A 2021, 824, 141842. [Google Scholar] [CrossRef]
  39. ASTM E10-15; Standard Test Method for Brinell Hardness of Metallic Materials. ASTM International: West Conshohocken, PA, USA, 2015.
  40. ASTM-E08; Standard Test Methods for Tension Testing of Metallic Materials. ASTM International: West Con-shohocken, PA, USA, 2022.
  41. Zhang, L.; Shi, G.; Xu, K.; Hao, W.; Li, Q.; Junyan, W.; Wang, Z. Phase Transformation and Mechanical Properties of B4C/Al Composites. J. Mater. Res. Technol. 2020, 9, 2116–2126. [Google Scholar] [CrossRef]
  42. Custodio, A.G.; Lindquist, K.J.; Tolentino, M.; Aranas, C.; Saha, G.C. Investigating the Nanoscale Hardness/Strength Properties of High-Entropy Alloy Particles Using the Nanoindentation Technique. J. Alloys Metall. Syst. 2023, 4, 100043. [Google Scholar] [CrossRef]
  43. Ma, X.; Cheng, B.; Zhao, K.; Yin, F.; Liu, X.; Chawla, N.; Zhao, Y. Microstructure Evolution and Mechanical Properties of High Fraction Al3BC/Al Composites Fabricated by a Reactive Hot Pressing Sintering Process. J. Alloys Compd. 2024, 1003, 175573. [Google Scholar] [CrossRef]
  44. Fjellstedt, J.; Jarfors, A.E.W. On the Precipitation of TiB2 in Aluminum Melts from the Reaction with KBF4 and K2TiF6. Mater. Sci. Eng. A 2005, 413–414, 527–532. [Google Scholar] [CrossRef]
  45. Bernacki, M. Kinetic Equations and Level-Set Approach for Simulating Solid-State Microstructure Evolutions at the Mesoscopic Scale: State of the Art, Limitations, and Prospects. Prog. Mater. Sci. 2024, 142, 101224. [Google Scholar] [CrossRef]
  46. Huang, K.; Marthinsen, K.; Zhao, Q.; Logé, R.E. The Double-Edge Effect of Second-Phase Particles on the Recrystallization Behaviour and Associated Mechanical Properties of Metallic Materials. Prog. Mater. Sci. 2018, 92, 284–359. [Google Scholar] [CrossRef]
  47. Sakai, T.; Belyakov, A.; Kaibyshev, R.; Miura, H.; Jonas, J.J. Dynamic and Post-Dynamic Recrystallization under Hot, Cold and Severe Plastic Deformation Conditions. Prog. Mater. Sci. 2014, 60, 130–207. [Google Scholar] [CrossRef]
  48. Gnjidić, Ž.; Božić, D.; Mitkov, M. The Influence of SiC Particles on the Compressive Properties of Metal Matrix Composites. Mater. Charact. 2001, 47, 129–138. [Google Scholar] [CrossRef]
  49. Sathish, S.; Venkatesh, J.; Johnson, P.; Annar, S.; Saranya, J.; Sai Chandra, J.; Subbiah, R.; Tharmalingam, S. Strain Hardening Behaviour of Friction Stir Welded Magnesium Alloy. Mater. Today Proc. 2023, 74, 34–39. [Google Scholar] [CrossRef]
  50. Luo, J.; Mei, Z.; Tian, W.; Wang, Z. Diminishing of Work Hardening in Electroformed Polycrystalline Copper with Nano-Sized and Uf-Sized Twins. Mater. Sci. Eng. A 2006, 441, 282–290. [Google Scholar] [CrossRef]
  51. SONG, M. Effects of Volume Fraction of SiC Particles on Mechanical Properties of SiC/Al Composites. Trans. Nonferrous Met. Soc. China 2009, 19, 1400–1404. [Google Scholar] [CrossRef]
  52. Shabani, A.; Toroghinejad, M.R.; Shafyei, A.; Logé, R.E. Evaluation of the Mechanical Properties of the Heat Treated FeCrCuMnNi High Entropy Alloy. Mater. Chem. Phys. 2019, 221, 68–77. [Google Scholar] [CrossRef]
  53. Zhang, Y.; Liu, J.P.; Chen, S.Y.; Xie, X.; Liaw, P.K.; Dahmen, K.A.; Qiao, J.W.; Wang, Y.L. Serration and Noise Behaviors in Materials. Prog. Mater. Sci. 2017, 90, 358–460. [Google Scholar] [CrossRef]
  54. Wei, S.; Moriarty, D.P.; Xu, M.; LeBeau, J.M.; Tasan, C.C. On the Plastic Deformation of a CoCrFeNiW-C Alloy at Elevated Temperatures: Part I. Serrated Plastic Flow and Its Latent Dynamics. Acta Mater. 2023, 242, 118430. [Google Scholar] [CrossRef]
  55. Hu, Q.; Guo, W.; Xiao, P.; Yao, J. First-Principles Investigation of Mechanical, Electronic, Dynamical, and Thermodynamic Properties of Al3BC. Phys. B Condens. Matter 2021, 616, 413127. [Google Scholar] [CrossRef]
  56. Yuan, Z.; Ma, Z.; Zhang, H.; Wang, K.; Yu, Y.; Li, S.; Zhang, X.; Wang, J.; Zhang, D. Macro-Micro Analysis of Mechanical Properties of Al0.6CoCrFeNi High-Entropy Alloy Particle-Reinforced Al-Based Composites. Mater. Sci. Eng. A 2023, 885, 145625. [Google Scholar] [CrossRef]
  57. Maity, T.; Prashanth, K.G.; Balci; Kim, J.T.; Schöberl, T.; Wang, Z.; Eckert, J. Influence of Severe Straining and Strain Rate on the Evolution of Dislocation Structures during Micro-/Nanoindentation in High Entropy Lamellar Eutectics. Int. J. Plast. 2018, 109, 121–136. [Google Scholar] [CrossRef]
  58. Kim, H.S. On the Rule of Mixtures for the Hardness of Particle Reinforced Composites. Mater. Sci. Eng. A 2000, 289, 30–33. [Google Scholar] [CrossRef]
  59. Tjong, S.C.; Ma, Z.Y. Microstructural and Mechanical Characteristics of in Situ Metal Matrix Composites. Mater. Sci. Eng. R Rep. 2000, 29, 49–113. [Google Scholar] [CrossRef]
  60. Chen, Y.; Ren, Y.; Li, K.; Dang, B.; Jian, Z. Laser Powder Bed Fusion of Oxidized Microscale SiC-Particle-Reinforced AlSi10Mg Matrix Composites: Microstructure, Porosity, and Mechanical Properties. Mater. Sci. Eng. A 2023, 870, 144860. [Google Scholar] [CrossRef]
  61. Cha, S.I.; Kim, K.T.; Arshad, S.N.; Mo, C.B.; Hong, S.H. Extraordinary Strengthening Effect of Carbon Nanotubes in Metal-Matrix Nanocomposites Processed by Molecular-Level Mixing. Adv. Mater. 2005, 17, 1377–1381. [Google Scholar] [CrossRef]
  62. Zhu, J.; Cai, Y.; Zhang, Y.; Liu, X.; Tian, J.; Ma, J.; Shen, J. Particle-Reinforced Cu Matrix Composites Fabricated by Sintering Core–Shell-Type Amorphous/Crystalline Composite Powders. Mater. Sci. Eng. A 2022, 854, 143823. [Google Scholar] [CrossRef]
  63. Martínez-Flores, E.; Negrete, J.; Villasenor, G.T. Structure and Properties of Zn–Al–Cu Alloy Reinforced with Alumina Particles. Mater. Des. 2003, 24, 281–286. [Google Scholar] [CrossRef]
  64. Pugh, S.F., XCII. Relations between the Elastic Moduli and the Plastic Properties of Polycrystalline Pure Metals. Lond. Edinb. Dublin Philos. Mag. J. Sci. 1954, 45, 823–843. [Google Scholar] [CrossRef]
  65. Surucu, G.; Yildiz, B.; Erkisi, A.; Wang, X.; Surucu, O. The Investigation of Electronic, Anisotropic Elastic and Lattice Dynamical Properties of MAB Phase Nanolaminated Ternary Borides: M2AlB2 (M=Mn, Fe and Co) under Spin Effects. J. Alloys Compd. 2020, 838, 155436. [Google Scholar] [CrossRef]
Figure 1. The schematic process flow chart employed for the preparation of the Al3BC/Al composite through the powder metallurgical processing route.
Figure 1. The schematic process flow chart employed for the preparation of the Al3BC/Al composite through the powder metallurgical processing route.
Applsci 15 05189 g001
Figure 2. (a) Differential scanning calorimetry traces showing peaks around 660 °C (exothermic) and 681 °C (endothermic) indicating Al melting and solid-solid reaction, leading to the formation of Al3BC. (b) X-ray diffraction patterns of (B+GNPs)/Al powders as a function of milling time showing the presence of single-phase fcc-Al peaks. (c) X-ray diffraction pattern of the as-fabricated Al3BC/Al composite showing the presence of the following phases: Al, Al3BC, and AlB2.
Figure 2. (a) Differential scanning calorimetry traces showing peaks around 660 °C (exothermic) and 681 °C (endothermic) indicating Al melting and solid-solid reaction, leading to the formation of Al3BC. (b) X-ray diffraction patterns of (B+GNPs)/Al powders as a function of milling time showing the presence of single-phase fcc-Al peaks. (c) X-ray diffraction pattern of the as-fabricated Al3BC/Al composite showing the presence of the following phases: Al, Al3BC, and AlB2.
Applsci 15 05189 g002
Figure 3. Scanning electron microscopy images showing the presence of micron-size Al3BC/AlB2 (dense grey) clusters in a fine grain fcc-Al (brighter) matrix in (a) low magnification and (b) high magnification. (c) Energy-dispersive spectroscopy elemental distribution maps of B, graphene (C), and Al, showing distribution of the phases, with some presence of O.
Figure 3. Scanning electron microscopy images showing the presence of micron-size Al3BC/AlB2 (dense grey) clusters in a fine grain fcc-Al (brighter) matrix in (a) low magnification and (b) high magnification. (c) Energy-dispersive spectroscopy elemental distribution maps of B, graphene (C), and Al, showing distribution of the phases, with some presence of O.
Applsci 15 05189 g003
Figure 4. Electron back-scattered diffraction images of corresponding scanning electron microscopy image shown in (a). (b) shows the phase fraction map, and (c) orientation map along with the inverse pole figure images.
Figure 4. Electron back-scattered diffraction images of corresponding scanning electron microscopy image shown in (a). (b) shows the phase fraction map, and (c) orientation map along with the inverse pole figure images.
Applsci 15 05189 g004
Figure 5. (a) Stress-strain curve shows a high yield strength of up to 108 MPa and ultimate compressive strength of up to 284 MPa, with plasticity reaching 17%. Inset: serrated flow potentially categorized to Type-B serrations. (b) Strain hardening behavior of the composite shows multiple strengthening mechanisms (insets showing the two-phase strain hardening behavior).
Figure 5. (a) Stress-strain curve shows a high yield strength of up to 108 MPa and ultimate compressive strength of up to 284 MPa, with plasticity reaching 17%. Inset: serrated flow potentially categorized to Type-B serrations. (b) Strain hardening behavior of the composite shows multiple strengthening mechanisms (insets showing the two-phase strain hardening behavior).
Applsci 15 05189 g005
Figure 6. Scanning electron microscopy images taken on the fracture surface of the composite after the compression test showed (a) the primary cracks traveling along the Al3BC network with a few secondary cracks penetrating the matrix. (b) The presence of cracked/broken Al3BC while strongly bonded FG-Al grains tightly surrounded by matrix (white circle). (c) The presence of dimples (red arrow) and tear ridges (yellow arrow), along with cleavage planes, shows the mixed mode of fracture. (d) Shielded dense and fine microcracks in the FG-Al matrix enable Al3BC/Al to efficiently dissipate the deformation energy (yellow arrow).
Figure 6. Scanning electron microscopy images taken on the fracture surface of the composite after the compression test showed (a) the primary cracks traveling along the Al3BC network with a few secondary cracks penetrating the matrix. (b) The presence of cracked/broken Al3BC while strongly bonded FG-Al grains tightly surrounded by matrix (white circle). (c) The presence of dimples (red arrow) and tear ridges (yellow arrow), along with cleavage planes, shows the mixed mode of fracture. (d) Shielded dense and fine microcracks in the FG-Al matrix enable Al3BC/Al to efficiently dissipate the deformation energy (yellow arrow).
Applsci 15 05189 g006
Table 1. The summary of phase formation along lattice parameters, a; wt./mass fraction (%), V f ; area fraction, A f ; Young’s modulus; and hardness of as-evolved phases.
Table 1. The summary of phase formation along lattice parameters, a; wt./mass fraction (%), V f ; area fraction, A f ; Young’s modulus; and hardness of as-evolved phases.
PhaseVf (wt.%)Lattice Parameter (Å)Hardness (GPa)Modulus (GPa)Af (%)
Al3BC65 ± 2(a) 3.488 ± 0.003
(c) 11.535 ± 0.003
22 ± 7326 [20]52 ± 4
fcc-Al25 ± 2(a) 4.098 ± 0.0013 ± 15231 ± 2
AlB210 ± 1(a) 2.978 ± 0.001
(c) 3.259 ± 0.001
2 ± 0.525017 ± 3
Table 2. Mechanical properties of the composites and their comparison with published reports.
Table 2. Mechanical properties of the composites and their comparison with published reports.
Alloyσy (MPa)σs (MPa)εs (%)H (GPa)E (GPa)σcRRef.
Al3BC/Al
Sintering + Hot pressing
108 284 1715 ± 5 212 1.635.6Present study
Ti/1TiB275085610-128 0.14-[12]
Ti/4TiB257668927-1030.20-[12]
Al/15Mg2Si
Sintering + Hot extrusion
1142007-750.75-[17]
Al/20SiC
Sintering + Hot extrusion
64.61277--1.00-[51]
Al/20Al2O3 SPS-96-----[10]
Al/AlNp RT179 251-14-0.40-[13]
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

Maity, T.; Prakash, A.; Roy, D.; Prashanth, K.G. In Situ Al3BC/Al Composite Fabricated via Solid-Solid Reaction: An Investigation on Microstructure and Mechanical Behavior. Appl. Sci. 2025, 15, 5189. https://doi.org/10.3390/app15095189

AMA Style

Maity T, Prakash A, Roy D, Prashanth KG. In Situ Al3BC/Al Composite Fabricated via Solid-Solid Reaction: An Investigation on Microstructure and Mechanical Behavior. Applied Sciences. 2025; 15(9):5189. https://doi.org/10.3390/app15095189

Chicago/Turabian Style

Maity, Tapabrata, Aditya Prakash, Debdas Roy, and Konda Gokuldoss Prashanth. 2025. "In Situ Al3BC/Al Composite Fabricated via Solid-Solid Reaction: An Investigation on Microstructure and Mechanical Behavior" Applied Sciences 15, no. 9: 5189. https://doi.org/10.3390/app15095189

APA Style

Maity, T., Prakash, A., Roy, D., & Prashanth, K. G. (2025). In Situ Al3BC/Al Composite Fabricated via Solid-Solid Reaction: An Investigation on Microstructure and Mechanical Behavior. Applied Sciences, 15(9), 5189. https://doi.org/10.3390/app15095189

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