Next Article in Journal
Gold Compounds Inhibit the Ca2+-ATPase Activity of Brain PMCA and Human Neuroblastoma SH-SY5Y Cells and Decrease Cell Viability
Next Article in Special Issue
Application of Ferromagnetic Microwires as Temperature Sensors in Measurements of Thermal Conductivity
Previous Article in Journal
The Effect of a Slow Strain Rate on the Stress Corrosion Resistance of Austenitic Stainless Steel Produced by the Wire Laser Additive Manufacturing Process
Previous Article in Special Issue
Effect of Oxygen Flow Rate on Properties of Aluminum-Doped Indium-Saving Indium Tin Oxide (ITO) Thin Films Sputtered on Preheated Glass Substrates
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Review of the Recent Development in Metallic Glass and Its Composites

Centre of Composite Materials, National University of Science and Technology «MISiS», Leninsky Prosp. 4, 119049 Moscow, Russia
*
Author to whom correspondence should be addressed.
Submission received: 22 September 2021 / Revised: 20 November 2021 / Accepted: 26 November 2021 / Published: 29 November 2021

Abstract

:
Metallic glasses are known for their mechanical properties but lack plasticity. This could be prevented by combining them with other materials or by inducing a second phase to form a composite. These composites have enhanced thermo-physical properties. The review paper aims to outline a summary of the current research done on metallic glass and its composites. A background in the history, properties, and their applications is discussed. Recent developments in biocompatible metallic glass composites, fiber-reinforced metallic glass, ex situ and in situ, are discussed.

1. Introduction

Metallic glass is known for its advanced engineering properties. They are amorphous. In 1960, Klement et al. from the California Institute of Technology developed the first metallic glass by using gold silicon alloy [1]. It was detected that 25 percent of the gold was amorphous when quenched. In 1966, studies were conducted on the effect of different cooling rates during rapid quenching of metals from spat cooling. This resulted in the formation of metastable phases, such as in amorphous alloys [2]. Binary Pd–Si glass, ternary Pd–Au–Si, and Pd–Cu–Si were prepared at room temperature by quenching at different cooling rates [3]. In 1975, Fe40Ni40B20 metallic glass was produced by a melt spinning technique. Inoue et al. showed the large glass-forming ability of the La–Al–Ni amorphous alloys [4]. Phase separation of Au55Pb22.5Sb22.5 metallic glass was studied and showed the influence of the surface energy on the metallic glass. It was concluded that surface energy had a contribution to the decomposition process [5]. Drehman et al. showed the preparation of Pd40Ni40P20 by the undercooling process. During formation, crystallization occurred uniformly, and the activation energy was calculated [6]. The Pd–Ge and Cu–Zr alloys had higher thermal properties, compared to metallic glass prepared by the melt-spinning technique [7]. Zr41.2Ti13.8Cu12.5Ni10.0 metallic glass was alloyed by the induction melting process at a cooling rate of less than 10 K/s. It had excellent glass-forming ability and the effect of crystallization was studied [8]. Metallic glass is mainly made up of two types: metal–metalloid and metal–metal combinations. In metal–metalloid [9], the metalloid constituent comprises 10–20% of the composition. The metal constituent is more than 80%. Some of the examples of metal–metalloid metallic glass are Pd77Cu6Si17, Fe40Ni40B20, and Ni49Fe29B6P14Si2. In metal–metal, the metal can have any compositional ratio. Ni60Nb40, La80Au20, and Fe90Zr10 are some examples. Some of the commonly used metallic glass are Zr-based [10,11,12,13], Fe-based [14,15,16], Co-based [17,18], Ni-based [19,20], Cu-based [21,22], and Pd-based [23,24] types. Metallic glass is used in biomedical applications [25,26]. Zr-based metallic glass can be used in stents [27]. Fe–B–Nd–Nb metallic glass is used in micro cantilevers [28]. Metallic glass can be used as a micro scanner [29]. They are used in electrical applications [30] and catalysis [31]. Metallic glass is generally prepared by the rapid cooling of the molten liquid [32], melt-spinning [7,33], powder metallurgy [34], liquid squat quenching [35], magnetron sputtering [36], and pulsed laser quenching [37]. Metallic glass is formed from the continuous cooling of the liquid state. It is generally non-crystalline in nature. The major criterion for the formation of metallic glass is its glass-forming ability (GFA). The glass formation ability is dependent on the critical cooling rate (Rc). The critical cooling rate (Rc) is the minimum cooling rate required to keep the molten amorphous, without any formation of the crystal precipitates during solidification. The lower the Rc, the higher the glass-forming ability (Table 1). The main criteria that are used to measure the glass-forming ability are reduced glass transition temperature (Trg) and supercooled region (Tsr). The reduced glass transition temperature (Trg) is defined as the proportion of the glass transition temperature (Tg) to the melting temperature (Tm) of the metallic glass. The supercooled region is the region in between the crystallization temperature (Tx) and the glass transition temperature (Tg). Metallic glass is formed near their eutectic points as compared to the non-eutectic points. The offset liquidus temperature (Tl) showed that Tg/Tl have a better correlation criterion to have a glass-forming ability than Tg/Tm [38]. The GFA is directly proportional to the thickness of the sample and inversely so to the critical cooling rate (Rc). Mukherjee et al. showed that the volume change in the crystallization is correlated with the viscosity at the melting temperature [39]. The fragility index (m) is related to the glass forming and plasticity of the metallic glass [40]. The glass forming ability was increased with the addition of Co in Ce70xAl10Cu20Cox. It was due to the increase in the Al-site symmetry and is a characteristic of the local structure.
Metallic glass is known for its mechanical properties [52,53] and has corrosion resistance [54]. The amorphous metals (Figure 1) have a better elastic limit and Young’s modulus, compared to (the mechanical properties of) the traditional alloys. The defects and grain boundaries in the crystalline solution make crystalline alloys susceptible to stress corrosion cracking and intergranular corrosion. Peter et al. showed the corrosion resistance of the metallic glass in an aqueous solution as compared to the crystalline materials [55]. They have higher wear resistance [56] and elastic behavior [57] compared to crystalline alloys.
Metallic glass exhibits physical–thermal properties in the supercooled region. They soften in the supercooled region and can be easily transformed into the desired form [58]. Mechanical properties of Zr–10Al–5Ti–17.9Cu–14.6Ni in the supercooled region had excellent mechanical properties. This was due to the formation of nano-crystallites in the supercooled region [59]. The Ti46Zr20V12Cu5Be17 metallic glass matrix composite showed an increase in the tensile ductility and necking in the supercooled region [60].
Metallic glass has no crystal defect, and has high strength, hardness, and yield strength. It has certain shortcomings, such as a brittle nature and low plasticity [61]. This is due to the absence of crystal defects and grain boundaries. The lack of ductility causes a deterioration in the mechanical properties of the metallic glass. The mechanical properties of the metallic glass in a non-equilibrium state are shifted toward the metastable equilibrium. This physical process is known as structural relaxation [62]. This causes modification in the physical properties, such as viscoelastic properties, elasticity, anelastic, magnetic, and corrosion, and other properties that are influenced by structural relaxation [63]. There are two kinds of structural relaxations: α-relaxation and ß-relaxation. α-relaxation causes glass transition phenomena and vitrification [64]. It generally disappears below glass transition temperature [64]. The beta-relaxation occurs at high temperatures, affects the mechanical properties of the metallic glass, and influences the plasticity mechanism [65]. The brittleness of the metallic glass can be overcome by the activation of shear zones [66]. It is defined as an irreversible shear strain that occurs when stress is applied to a localized deformed non-crystalline solid. This, in other words, can be called the formation of the shear bands [67]. When stress is induced, plastic deformation occurs by the shearing of the material relative to its resting part. This causes a change in the shape of the material. The localized cluster is plastically deformed, causing it to become softer than its undeformed surroundings. This causes more flow of plastic strain into thin bands. They can occur by external and internal factors. Maass et al. showed that the strength of Zr52.2Ti5Cu17.9Ni14.6Al10 increases with a decrease in temperature. Shear band activity is increased with an inhomogeneous flow through a single shear band at cryogenic temperatures [68]. The shear band density, bending strain, and other mechanical properties of Pt57.5Cu14.7Ni5.3P22.5 and Zr35Ti30Cu8.25Be26.75 metallic glasses were decreased with an increase in the temperature. When the same strain was applied, the shear band at lower temperature produced multiple bands with smaller lengths, leading to increased properties [69].
Due to the catastrophic failure and brittle nature of metallic glass, researchers have developed composites with excellent thermo-physical properties.
The aim of the present work is to provide a brief overview of the recent development in metallic glass with a stress on the progress in the last 10 years. Please note that bulk metallic glass is considered and written as metallic glass in this review.

2. Biocompatible Metallic Glass and Their Composites

Metallic glasses are new in the field of biomedical applications with superior toughness, strength, and ductility, compared to crystalline alloys. Earlier, metallic glasses were used for structural engineering, due to their high mechanical properties. Metallic glasses are nowadays used as biomaterial implants.
Metallic crystalline materials, such as titanium alloys, Zr alloys, and stainless steel, are used in cardiovascular stents, hip joints, and bone implants (orthopedic joints) [70].
These crystalline metals have low strength, low wear, and corrosion resistance. Complex implants can be made from metallic glass because of their atomic structure, promising chemistry, and surface configurations. These properties can construct a biological response (Figure 2) with excellent properties over the present metallic biomaterials. Figure 2 also explains that the combination of amorphous structures with metallic constituents produces bio-metallic glasses. Biomedical metallic glasses have a high elastic modulus at 2%, compared to human bone (1%) and under stress, they have a unique ability to flex with the natural bending bone. This concludes that stress will be uniformly distributed, reducing the stress shielding effect, and the healing rate will be achieved faster [71].
Metallic glass is used in medical–surgical equipment [72], cardiovascular stents [73] and orthopedic implants [74].

2.1. Ti-Based Biomaterial Metallic Glass

Titanium-based metallic glass is used in implant applications and has better corrosion properties, higher fracture strength, and a lower Young’s modulus than crystalline materials. Ti45Zr10Cu31Pd10Sn4 [75] was prepared by gas atomization and porous Ti45Zr10Cu31Pd10Sn4 by Spark Plasma sintering (SPS). Porous metallic glass has better functionality than the commercially used Ti and Ti–6Al–4V implantable biomaterials. The decrease in the porosity in the metallic glasses causes an increase in Young’s modulus and the compressive strength and a decrease in the corrosion resistance. It showed better biocompatibility than the commercial implantable biomedical materials. Similar work was done by Xie et al. [76] in which Ti45Zr10Cu31Pd10Sn4 metallic glass was made by the combination of gas atomization and NaCl powder and SPS. NaCl powder was used to keep the porosity under control. The composite exhibited three times higher yield strength than human bone and could be used for surgical implants. Ti51Zr5Cu41Sn3 [77] showed better wear resistance and tribological properties than the pure Ti in the SBF (simulated body fluid). It also showed good biocompatibility in vitro than theTi45Zr5Cu41Ni9 and Ti45Zr5Cu41Ni6Sn3. A thin-film metallic glass [78] of different ratios, Ti–Cu–Pd–Zr and Ti–Cu–Pd–Zr: B were fabricated by pulsed laser deposition (PLD). Boron-added metallic glass (Ti–Cu–Pd–Zr: B) showed good biocompatibility and corrosion resistance. The glass transition temperature was increased with an increase in the boron percentage. There was non-agglomeration of the RBC (red blood cell) in TFMGs after 30 minutes. The morphology of the TiB-2 cell (Figure 3) shows the disturbed nature with the addition of EDTA to the cell.
Ti-based metallic glasses Ti45Zr40Si15 (Cu-free), Ti45Zr40Si10Cu5 (low-Cu), and Ti45Zr20Cu35 (high-Cu) [79] were prepared by different Cu proportions. It showed that metallic glass with low content of Cu had good biocompatibility and low cytotoxicity, compared to pure Ti. Ti45Zr20Cu35 (high Cu) showed poor in vitro viability, due to the higher content of the Cu. This suggests that the Ti45Zr40Si15 (Cu-free), can be used in biomaterials. Ti47Cu38−xZr7.5Fe2.5Sn2Si1Ag2Tax (x = 1–4) [80] were prepared by copper mold casting. The supercooled region of the metallic glass decreased with an increase in Ta content. The material showed good bio-corrosion resistance, mechanical properties, and a large, supercooled region. Laser surface remelting was used from an in situ bonded coating of the Ti47Cu38Zr7.5Fe2.5Sn2Si1Nb2 metallic glass [81]. The additional Nb formed an oxide layer, which prevented corrosion. There was an increase in the hardness and a decrease in the elastic modulus. The non-toxic nature, low elastic modulus, and corrosion resistance (Table 2) can ensure the using of metallic glass as a biocorrosion biomaterial. Nb was added to Ti–Cu–Zr–Fe–Sn–Si–Ag [82] metallic glass, and Ti47Cu38−xZr7.5Fe2.5Sn2Si1Ag2Nbx (x = 0, 1,2; at%) was prepared. The addition of Nb showed a decrease in the glass transition temperature, superior bio-corrosion resistance, good in vitro compatibility, and excellent mechanical properties. Ti40Zr10Cu36Pd14 [83] metallic glass was compared with conventional Ti–6Al–4V and could be used in dental implantology. It showed high thermal stability, cytocompatibility, resistance to sterilization, and good corrosion resistance. These properties ensure that the material can be considered biomaterial. A glassy matrix Ti75Zr10Si15 with embedded Ti60Zr10Nb15Si15 [84] was prepared by melt spinning with the formation of single ß-type nanocrystals. Stable surface passivity was due to the homogeneous distribution of the constituents. Nb improved the glass-forming ability, mechanical properties, and resistance toward the pitting corrosion. Hydroxyapatite was formed in the metallic glass in SBF (simulated body fluid). Ti60Zr10Nb15Si15 showed wear and bio-corrosion resistance. The oxide layer was formed in Ti41.5Zr2.5Hf5Cu37.5Ni7.5Si1Sn5 [85] and showed good corrosion resistance, and hardness. Pitting corrosion was higher in the artificial saliva solution than in SBF (simulated body fluid). Ti41.5Zr2.5Hf5Cu37.5Ni7.5Si1Sn5 has lower cell viability L929 and NIH3T3 cells. This shows that metallic glass can be used in bone tissue and has excellent osseointegration. Ti-based nano glass composite with sub-micro-nanometer-sized hierarchical glassy structure was prepared by magnetron sputtering. Enhancement in cell proliferation, better biocompatibility, and bioactivity was attained by the formation of button-like clumps [86].

2.2. Zr-Based Biomaterial Metallic Glass

Zr-based metallic glass has anticipated properties in orthopedic applications [87]. It is used in arthroplasty, bone screws, and intramedullary nails. The main problem with Zr-based metallic glass is that it does not allow adhesion and tissue growth at the host site. Metallic glass Zr44Ti11Cu10Ni11Be25 [88] was converted into ceramic films. Ceramic conversion treatment (CCT) in an oxidizing medium was used in the supercooled region of the metallic glass to convert its surface. Thermal oxidation was performed at 350 °C for 40 h and at 380 °C for 4.5 h to alter the surface properties. There was an increase in the surface hardness and Young’s modulus (Table 3), and the coefficient of friction was reduced. SAOS-2 human osteoblast-like cells increased surface coverage for the untreated surface. This shows that metallic glass increased its biocompatibility and tribological properties. The surface quality of the Zr-based metallic glass (Zr67Cu10.6Ni9.8Ti8.8Be3.8) [89] was improved by bioactive hydroxyapatite suspended powder in dielectric fluid from electro-discharge machining (EDM). The surface hardness was increased and a crack-free HA-EDMed surface, with nanopores, was achieved. Tribocorosion was studied for Zr65Cu18Ni7Al10, and Zr55Cu30Ni5Al10 found that Zr65Cu18Ni7Al10 [90] has better plasticity and can be used as a potential load-bearing implant. Abrasive wear was experienced in both metallic glasses. The passive nature of the Zr65Cu18Ni7Al10 showed a negative effect on tribological corrosion. Zr48Cu36Al8Ag8 [91] thin film (at.%) was prepared by a 316L stainless steel substrate by applying magnetron sputtering. The membrane should have good scratch resistance, mechanical properties, and lower Young’s modulus. TFMG-coated steels showed good corrosion resistance in SBF (simulated body fluid) and a non-cytotoxic nature in L929 cells. Low energy ion implantation [92] was done on the (Zr0.55Al0.10Ni0.05Cu0.30)99Y1 to improve the biocompatibility and activity of the metallic glass. Ca-ion showed an increase in cell adhesion and changed the cell behavior. A higher cell proliferation was observed in Ni-free Zr60 + xTi2.5Al10Fe12.5−xCu10Ag5 (at.%, x = 0, 2.5, 5) [93]. Zr55Cu30Ni5Al10 based metallic glass shows the highest wear rates in the different electrolyte systems (NaCl, phosphate solution with and without protein) as compared to the crystalline structure. The formation of the nanocrystalline (Zr2Cu) leads to the formation of a passive layer, leading to corrosion resistance (Figure 4). The author concluded that the Zr-based metallic glass should be mixed with the (Zr2Cu) nanocrystalline particles to improve the passivation [94].
Ni-free Zr65Ti2.5Al10Fe7.5Cu10Ag showed good mechanical properties and in vitro biocompatibility [95]. Zr55Al10Cu30Ni5 was alloyed with 1 at.% yttrium for surface roughness and cell behavior. The addition of 1 at.% yttrium did not have any substantial effect on proliferation and cell adhesion, but a decrease in alkaline phosphatase (ALP) activity was seen [96].

2.3. Magnesium-Based Biomedical Metallic Glass

Magnesium alloys degrade over time and do not need surgery to remove them. They have a density similar to bone (1.8–2.1 g/cm3) and can be used as an allograft [97]. In vivo studies of Mg66Zn30Ca4 and PCL composite [97] were examined and it was found that the composite was biocompatible. The composite properties can withstand 100% elongation when stress is applied and have thermal properties. Scaffolds of PCL/Mg metallic glass and PCL/nHA/Mg metallic glass [98], when immersed in SBF (simulated body fluid), showed their bioactivity. The mechanical properties were affected faster in the early stages of the hydrolysis but were slower later regarding in vitro loss. The composite scaffolds also controlled magnesium release. The scaffolds have better mechanical properties, hemocompatibility and can be used in orthopedic implants. The bio-corrosion and cytotoxicity of two relaxed and crystallized Mg70Zn26Ca4 ribbons [99] in MTT (tetrazolium-based colorimetric assay) showed that the relaxed metallic glass had better biocorrosion resistance and biocompatibility. MTT (tetrazolium-based colorimetric assay) assay used with Schwann cells concluded its application in nerve tissue regeneration. Tribological studies of Mg65Zn30Ca5 [100] in phosphate-buffered saline (PBS) applied on ball-on-disk reciprocating sliding showed that the metallic glass had the highest wear resistance with a comparison with AZ31B alloy and pure Mg. The corrosion potential decreases the tribological contact and increases the corrosion current densities of the Mg-based alloys. The wear resistance suggests that magnesium-based metallic glass can be used in future orthopedic implants. A new composition of metallic glass Mg(85 − x)Ca(8 + x)Au7 (with x = 0, 2, 4) [101] and Mg(81−x)Ca10Au7Yb(2 + x) (with x = 0, 8) showed resistance to the crystallization and has the potential to be used in sterilization procedure in biomedical applications. The absence of non-toxic elements in the metallic also plays a major role in its biomedical usage. In vitro studies of Mg66Zn30Ca4-xSrx [102] (x = 0, 1 and 1.5 at.%) metallic glass was studied for the bone-forming MC3T3-E1 pre-osteoblast and for the viability in the orthopedic implant. The results showed that metallic glass had more cell adhesion compared to AZ31B alloy, caused the proliferation of pre-osteoblasts, and no cytotoxicity occurred. These properties suggest that metallic glass is biocompatible and has the potential for biomedical applications. Mg66Zn30-xCa4Agx [103] (x = 0, 1 and 3 at.%) showed that the addition of Ag improved corrosion and suppressed hydrogen evolution. Properties like corrosion resistance, and Ag alloy’s anti-bacterial and cytotoxicity properties make it biocompatible. The Mg-based metallic glass can be used as a biomaterial material. Comparison studies (Figure 5) were performed between Mg66Zn30Ca4xSrx [104] (x = 0, 0.5, 1 and 1.5 at.%) and MgZnCaSr. It showed that Mg66Zn30Ca4xSrx has the better glass-forming ability, increased corrosion resistance, good mechanical properties, and cytocompatibility. These properties show future potential in biomedical applications.
The alloying of 2% to 4% of Yb to Mg66Zn30Ca4 [105] increased the ductility of the composite. This is due to the presence of extensive shear bands in Mg66Zn30Ca2Yb and (Yb2) Mg66Zn30Yb4 (Yb4). Further, in vitro studies showed that the addition of 4% Yb decreases cytotoxicity. It was concluded that the Mg66Zn30Yb4 amorphous composite can be used in future biomaterial implants. Alloying Mn to Mg69−xZn27Ca4Mnx (x = 0, 0.5 and 1 at.%) showed a decrease in the mechanical strength and glass-forming ability of the composite. There was an increase in the bio-corrosion resistance (Table 4) and cell viability of osteoblasts. The strength of the composite is similar to bone implants and can be used as biomaterials. Metallic glasses with different thicknesses [106], Mg67Zn27Ca4 at different cooling rates (30 m/s (M1) and 10 m/s (M2)) were prepared by Mg67Zn28Ca5 metallic glass. It showed that metallic glass with a lower cooling rate had a lesser corrosion rate in the saline solution and was susceptible to pitting corrosion.

2.4. Fe- and Sr-Based Metallic Glass Biomedical Metallic Glass

Fe-based biomaterials are known for their load-bearing implants and are used in hip prostheses, dental implants, and bone fixators in the hip and knee. A thin layer of metallic glass (Fe37Cr15Mo2B26C7Nb3Si3Al6Mn1 (%at)) was [107] formed with a 316L stainless steel substrate by using the ESD method. The layer was biocompatible, uniform, dense, and showed high adhesion. J-quenching and flux treatment were used to prepared a Fe80 − x − yCrxMoyP13C7 (x = 10, y = 10; x = 20, y = 5; x = 20, y = 10, all in at.%) [108]. This metallic glass has higher corrosion resistance than the 316L stainless steel and TC4 and is biocompatible. Micrographs also showed that the Fe-based metallic glass was filled with NIH 3T3 cells, shown on the left side of Figure 6. This shows the biocompatibility of Fe-based metallic glass. Fe73.5Nb3Cu1Si13.5B9 [109] was ball milled for 48 h and three kinds of magnetic fluid, Fe73.5Nb3Cu1Si13.5B9MR fluid, CoFe2O4 FF, and Fe3O4 FF (ferrofluid), were prepared. The metallic glass showed the potential for use in hyperthermia in tumor therapy. The addition of Nb in Fe55−xCr18Mo7B16C4Nbx (x = 0, 3, and 4 at.%) [110] improved the corrosion resistance in the ringer solution and can be used as a biomaterial. Sr60Mg18Zn22-based metallic glass has a low glass transition temperature and has homogenous flow in the supercooled region. The chemical and physical properties of Sr60Mg18Zn22 suggest that it can be used as a biomaterial [111].

2.5. Summary

Bio-metallic glass materials and composites are known for their high mechanical properties and corrosion resistance. Zr-based metallic glasses have adhesion and host site limitations. Mg-based metallic glasses cause hydrogen gas to release and can cause degradation of the material. This can be prevented by adding ceramic coatings and metallic materials to the metallic glass matrix. It was shown that metallic glasses show better properties than traditional alloys.

3. Fiber-Reinforced Metallic Glasses Composites

Fiber-reinforced metal matrix composites are used for their mechanical properties and industrial applications, due to their low-cost production and performance. They are used in the aerospace and automobile industries. The formation of the oxide or carbide layer on the fiber can degrade its physical properties [112]. Metallic glass is known for its mechanical properties but lacks plasticity; a combination of fiber-reinforced metallic glass can produce a composite with good mechanical properties. In recent research, the vacuum pressure infiltration processing method was used to produce a carbon fiber/Ti37.3Zr22.7Be25.5Fe5.5Cu9 (at.%) composite [113]. Microstructure control, good bonding, and strain percentage were increased, due to the interaction between carbon fiber and shear bands in the metallic glass. Mechanical strength (Table 5) of the tungsten fiber/Zr41.2Ti13.8Cu10Ni12.5Be22.5 [114] was increased. This can be attributed to an increase in fiber volume fraction (Vf). Similar studies of tungsten fiber/(Zr40.08Ti13.30Cu11.84Ni10.07Be24.71)99Nb1 [115] increased the mechanical properties of the composite. Nb was added to increase the properties and interference of the composite. The yield stress of the composite gradually decreased with the shear softening. TiNi fibers [116] in metallic glass matrix generated the propagation of multiple shear bands and induced the TRIP effect, resulting in superior mechanical properties. The more the fiber volume fraction (Vf), the better the mechanical properties of the composite. Tungsten fiber [117] of various dimensions was reinforced in Zr-based metallic glass. The mechanical properties of the composites increase with a decrease in the size of the tungsten fiber. Inhibition of shear bands and different fracture modes caused the superior compressive properties. The maximum compressive strength for the composite was 3079 MPa. A composite was produced by a different volume fraction of the tungsten strings/Zr-based metallic glass [118]. The 68% volume fraction tungsten composite showed the highest mechanical properties among the different composites. The enhanced mechanical properties are attributed to the spatial geometry, which modifies the stress pattern, resulting in the formation of multiple shear bands in the composite. A laminate structure of carbon fiber-reinforced epoxy and Zr44Ti11Ni10Cu10Be25 showed an increase in the adhesion of the composite [119]. A composite [120] based on volume fraction tungsten fiber reinforced with metallic glass at different temperatures showed that the higher the volume fraction of the tungsten, the better the mechanical properties. Both the quasistatic strength and the dynamic strength decrease with an increase in the temperature. Mg-based metallic glass reinforced with carbon fiber has engineering applications [121]. Al matrix composites reinforced with Fe74Mo4P10C7.5B2.5Si2 (at.%) [122] metallic glass fibers were prepared by ball milling and hot pressing. The plastic deformation and strength were increased due to the presence of shear slip planes and the axial orientation of the fiber. The addition of the Nb in the W/Zr-based metallic glass composite can increase the interaction between the composite [123]. Chen et al. showed that for W/Zr-based composite, splitting of the composite and the critical energy dissipations for the shear banding are obtained as a function of the strain rate and tungsten volume fiber fraction [124]. It was found that the transition mode in the failure in the W/Zr composite is controlled by the shear to normal strength ratio [125]. B. Zhang et al. studied the effect of tungsten fiber orientation in a tungsten/Zr-based metallic glass composite. It was concluded that with the increase in orientation angle θf, the tensile strength decreased and no plasticity was present [126]. Deformation in W/Zr-based metallic glass occurs in two phases. In uniaxial compression, the metallic glass matrix deforms homogenously, but buckling of the tungsten fiber occurs first. After further deformation, shear band propagation occurs in the composite, and the splitting of tungsten fiber with shear bands takes place in the fracture of the composite [127]. A comparison between different composite W/Zr and W/Cu metallic glass showed that the W/Cu-based composite has higher cohesion strength and compression stress and strain. The fracture mechanism of the composites was related to the distinct glass-forming ability of the monolithic metallic glasses [128]. Shear bands in the 61.4 vol% (Figure 7) tungsten wires/ Zr-based metallic glass cause good mechanical properties and cohesion in the composite [129]. Wang et al. [130] showed that at a strain rate below 3000 s−1, the compressive strength of the composite Wf/Zr based metallic glass (60%) increased. A ballistic test on the W/Zr-based metallic glass composite showed that the penetration depth of the composite is 50% more than the tungsten heavy alloy rods (95W rod) [131]. The mechanical properties of the Wf/Zr-based metallic glass composite are increased with an increase in the Nb content. Nb increases the interfacial strength in between the tungsten fiber and metallic glass [132]. Stainless steel capillaries reinforced in Zr-based metallic glass showed an increase in plastic strain by 14% [133].

Summary

Fiberglass and metallic glass composites are low-cost in production. Tungsten b fiberglass reinforced with metallic glass produces high yield strength and plastic strain. The addition of Nb increases the interfacial strength between the metallic glass and fiber. The free volume also determines the properties of the fiber glass and metallic glass composites.

4. Metallic Glass/Polymer Composites

A high magnetoelectric voltage coefficient showed that the P(VDF-TrFE) polymer film laminated on metallic glass (Fe67Co18B14Si1) with an epoxy resin adhesive composite can be used for magnetoelectric energy harvesting [134]. Mg-based metallic glass/PCL composite shows biodegradable properties and can be used in biomedical applications [97]. Al-based metallic glass/PET composite was prepared by ball milling and subsequent spark plasma sintering. The author concluded that a change in the chemical composition took place due to the partial crystallization of the composite [135]. Similar work was done by producing a metallic glass/polymer composite by mechanical alloying and sparks plasma sintering. The composite showed that it can be used as an antifriction material and in tribology [136]. M.Y. Zadorozhnyy et al. found that the adhesion of the metallic glass/polymer composite can be increased by adding 10 mass% of triethoxyvinylsilane. The composite (HDPE/Mg67,5Ca5Zn27,5) also showed high thermal conductivity and elastic modulus [137]. A composite based on Zr65Cu17.5Ni10Al7.5 and PTFE polymer showed the formation of Zr2Cu and Zr2Ni intermetallic, and the addition of metallic glass increased the thermal conductivity in the composite [138]. A silicone polymer was used as a binder to form a metallic glass–polymer nanocrystalline structure (Fe-based composite). It was also concluded that the composite showed improvement in soft magnetic properties [139]. Zr-based metallic glass/polyacrylonitrile (PAN) composite can be used to produce nitrogen selective applications [140]. Metallic glass–polyisoprene nanolaminate causes an increase in the plastic flow and can be used in structural and nanodevices. The composite also shows corrosion and wear resistance [141]. A similar study was conducted by producing Fe64Co17Si7B12 metallic glass and polyvinylidene fluoride (PVDF) using an epoxy resin, which has a piezoelectric response and can be used in miniature electronic devices and ultra-low power applications [142].

Summary

Metallic glasses have high strength but lack plasticity. The polymer has low strength and high plasticity. The combination of both can produce a composite with good thermo-physical properties. The lack of interaction between the polymer and metallic glass is a major drawback and can be avoided by using binders. M.Y. Zadorozhnyy [137] showed that silane can be used as a binder between metallic glass and polymer.

5. Ex Situ Metallic Glass Composites

Ex situ [143] metallic glass composites are generally formed by mechanical alloying or atomization of the powder. These composites have a glassy phase and reinforcement of second phase particles. The variation in particle size varies from micrometer to nanometer. It can also be formed by introducing a second phase as reinforcement in the glassy matrix. Metallic glass matrix/fiber reinforcement composite are some of its examples. A Ti-based metallic glass composite showed that with an increase in the Zr content, the mechanical properties improved. The solid solution strengthening of ex situ Ti layers hinders the propagation of shear bands and improves the mechanical properties [144]. The addition of Nb in Ti-based metallic glass increases the bond interference and mechanical properties (Table 6) of the composite. The formation of the α-Ti gradient layer helps in restraining the shears bands, which leads to the formation of multiple shear bands and increases the plasticity of the composite [145]. The electroless copper plating forms a stable and protective layer between SiC particles and Mg-based metallic glass. The SiC particles increase the plastic zone size, and the Cu coating helps in the prevention of shear bands, improving the plasticity [146]. The ex situ Mg-based metallic glass composite showed that shear band stability can hinder the motion of shear bands [147]. A work hardening behavior was seen in the composite because of the porous NiTi particle in Mg-based metallic glass. The hindrance of the movement of the single shear bands leads to the increase in the plasticity of the composite [148]. Zr-based metallic glass drill bits can be used in orthopedic drills in medical applications [149]. Ex situ Ta (0–6 vol%) particles were reinforced by pulse laser technique on Zr-based metallic glass, and no harmful crystalline particles were present. There was the formation of sub-zero and nano-size Ta particles in the weld fusion zone. Ta also affected the glass transition temperature and glass-forming ability of the composite [150]. An interfacial reaction occurred when an ex situ composite was welded. There was the formation of an interfacial layer in the reinforced phase (Ta) in PM, WFZ, and HAZ zones. A more stable ZrCu layer was formed in WFZ [151]. The La-based composite showed good mechanical properties but can cause plastic strain, due to the formation of oxide defects [152]. Ti particles in Mg-based composite increase the plasticity of the composite due to the multiple shear bands. The plasticity of the composite depends on Ti particle size and interparticle spacing [153]. In similar studies, J.B. Li et al. showed that shear band propagation (Figure 8) was restricted by the semi-uniform confinement zones of Ta particles. The plasticity was dependent on the interparticle size and confinement zone [154].
Ex situ Ta particles in the Zr-based composite showed an increase in plasticity due to shorter interplanar spacing. The interfacial area was also increased due to the bigger volume fraction particles [155].

6. In Situ Metallic Glass Composites

The plasticity of the metallic glass composite can be increased by introducing a second phase, which can block the propagation of shear bands. The formation of multiple shear bands can increase the plasticity of the metallic glass composite. The crystalline phases occur during the solidification of the melts, which increases the plasticity. These composites are in the form of crystalline phases, nanocrystalline phases, and dendritic phases [156]. Ex situ metallic composites are generally prepared by techniques such as ball milling and the melt infiltration process [142]. Different particles of size ranging from micrometer to nanometer are induced. In situ composites are produced by thermal treatment, high-pressure torsion, and hot extrusion processes. Secondary particles, such as crystalline, semicrystalline, and quasi-crystalline, are formed. By controlling the kinetics and thermodynamics of the metallic glass in the supercooled region, the crystallization process can be avoided. The heat treatment process is generally done to control the crystallization process.

6.1. Crystalline/Nanocrystalline Metallic Glass Composites

A dual crystalline phase, α- and β-phase reinforced in Ti-based metallic glass composite showed good mechanical properties. This could be attributed to the hindrance of shear bands by crystalline phases and α-phase toughening in the β-phase [157]. The high torsion method was used to prepare a nanocrystalline composite and a Cu-rich nanocrystal increased the yield and ultimate strength of the metallic glass. The strain hardening and elongation were increased by the increase in the free volume [158]. Nanoporous metallic glass (Fe–Ni–Co) composite can be used in clean energy applications [159]. Martensitic transformation of nanocrystalline Ti50Ni47Fe3 composite can provide near-ideal strain and strength in a composite [160]. FeCrMoCB metallic glass can be used in joint replacement implants [161]. Uniformity in the eutectic phase and Fe2B nanocrystalline structure plays an important role in the properties of a material. The nanocrystalline structure of Fe-based metallic glass enhances the mechanical properties (Table 7) of the composite [162]. The enhancements in the absorption kinetics of Mg-based metallic glass occur due to the formation of nanocrystals by the control agents present in the ball milling process. The dehydrogenation in the Mg60Ce10Ni20Cu10 metallic glass was controlled by the formation of nanocrystal hydrides [163]. Nanocrystals in the Fe-based metallic glass that occurred during partial annealing led to an increase in the ferromagnetic properties of the composite. Crystalline particles present in the composite increased the volume fraction, which resulted in increasing the hardness of the composite [164]. Q. Zhou et al. showed that the addition of Sn to Ti-based metallic glass has wear resistance and can produce the formation of a continuous tribolayer. This tribolayer during sliding shows improved the tribological properties [165]. A. Yao et al. showed that the nanocrystalline metallic glass (Al82Ni6Co3Y6Cu3) can be used as an easy precursor in manufacturing non-porous electrodes [166]. Crystalline FexN in FeSiCrB metallic ribbon provided high electric resistance, magnetic moment, works as a binder, and boosts magnetic properties in the FeSiCrB metallic ribbon. The stacking factor of the magnetic material is near 100% and the composite can be used as a soft magnetic material in electrical and electronics machines [167]. The composite of Zr/nanocrystalline Fe metallic glass has better compressive strength and hardness than the Zr metallic glass [168], but 40% above the volume fraction of Fe-based metallic glass causes cracks and pores formation. X. Ji et al. discussed that microstructure evolution is related to the solidification of the cooling rate, impact-induced temperature rise, plastic flow, oxidation, and interdiffusion of the metal elements [169]. Non-planar thin film (CuO-Cu2O)/ZnO: Al heterojunctions have applications in solar and electronic devices [170]. W. R. Jian et al. studied that the volume fraction of the Cu64Zr36 crystal particles can achieve a balance between hardness, hardness, and ductility [171]. Partial nano-crystallization of Zr-based composite has anticipated vibration and electric transport properties [172]. Microstructure analysis (Figure 9) of the Ti-based composite showed that the crystallization occurred in the amorphous phase in two phases, and the formation of nanocrystals (B2) and precipitation of the Ɣ-TiCu phase occurred afterward [173].
Cu-based metal–intermetallic nanostructured composites have high thermal stability, high ductility, and superior strength because of their distinctive microstructure and irregular deformation mechanism. This composite can have applications in construction, military, automobile, and aerospace [174].
The formation of precipitation of the nano-sized Ni2SnTi phase occurs during annealing in the metallic glass (Ti45Cu41Ni6Zr5Sn3), and an increase in the wear resistance, microhardness, and wear resistance is induced by the nanocrystals [175]. J.-K. Lee et al. showed that the deformation of W/Ni-based metallic glass composites depends on the layered structure of the nanograined tungsten matrix and its reinforcement [176]. The addition of boron in the Mo–Ni–Si–B metallic glass can increase the glass-forming ability of the composite. Mo–Ni–Si–B metallic glass can be used to form a metallic glass–nitride nanocomposite coating layer and has an application in wear resistance coatings [177]. The porosity structure of Zr70Cu24Al4Nb2 metallic glass can be influenced by the SPS procedure. The higher sintering temperatures lead to the formation of nanocrystals in the composite. This composite has an application in functional materials [178]. The coercivity of the Co72B19.2Si4.8Cr4 metallic glass is changed by the intermediate nanocrystalline phase, resulting in an amorphous soft state to a monocrotaline hard magnetic state [179]. Q. Xing et al. showed that pre-compression treatment for RE65Co25Al10 metallic glass can induce the formation of the nanocrystalline composite. The pre-compression treatment is good for thermal and mechanical properties [180]. The decrease in the plasticity of Ti-based metallic glass is due to the decimation of the free volume and microcracks, and precipitation of nanocrystals leads to brittle fracture [181]. R. Babilas et al. showed that in Mg60Cu30Y10, the homogenous metallic glass has more corrosion resistance than the nanocrystalline particles in the composite [182]. The nanocrystalline Mg rod obstructs the propagation of shear bands in thin-film metallic glass. The more the Mg content in the thin film, the better the ductile behavior. The author concluded that the composite can be used in micro-electronic–mechanical biomedical devices [183]. The increase in the fraction ß-type phase leads to the degradation and decrease in the mechanical properties in Ti60Zr10Nb15Si15 as compared to Ti75Zr10Si15. This leads to a high H/ER ratio and proposes an application in high wear resistance [184]. Ag in Zr56Co22Al16Ag6, in the local structure of the Zr–Co–Al, is influenced by the positive mixing of Ag and Co in Zr56Co22Al16Ag6. The nucleation rate and growth rate at a maximum temperature are credited to ultrafine nanocrystals [185]. L.M. Zou et al. showed that the addition in the content of Fe in TiNbZrTa increased the glass-forming ability and enthalpy of crystallization and transformed the structure from a nanocrystalline structure in the nanocomposite matrix with nanocrystals in an amorphous matrix [186]. A sintered and porous structure Ti-based metallic glassy/nanocrystalline composite has application in structural and functional applications [187]. The increase in the Cu–Ni content in Ti–Zr–Cu–Ni–Co amorphous alloys showed an increase in the compressive strength and plasticity in the composite [188,189]. Dynamic martensitic transformation and heterophase structure increased the strength and plasticity of the composite in the Ni–Cu–Ti–Zr composite [190]. The addition of Fe in the Ti–Cu amorphous alloy led to the formation of CuTi2 intermetallic compounds, which degraded the mechanical properties of the composite [191]. I.S Golovin et al. studied the change in the internal friction values; the structural changes occurred due to the martensitic transformation in the Ni40Cu10Ti33Zr17 composite [192]. The addition of Y, Nb, and Al in (Ti–Ni–Cu–Zr) metallic glass induced transformational plasticity and an increase in yield strength [193]. Structural relaxation occurred in the Zr–Cu–Ni–Al metallic glass, storage modulus and internal friction changed due to the amorphous alloy crystallization [194]. The formation of the NiTi2 phase in the (Ti–Ni)–(Cu–Zr) phase caused degradation in the mechanical properties of the composite [195].
Table 7. Mechanical properties of crystalline/nanocrystalline metallic glass composites.
Table 7. Mechanical properties of crystalline/nanocrystalline metallic glass composites.
Chemical CompositionPreparation MethodYield Strength (MPa)Ultimate TensilePlastic StrainReference
Ti31.2Zr28.7Ni5Cu8.5Be21.6Mo5Melt Spinning/quenching141016253.6%[157]
Zr52.5Cu17.9Al10Ni14.6Ti5High-Pressure Torsion179320230.4%[158]
Ti50Ni47Fe3Arc melting/annealing170022505.5%[160]
Zr48Cu36Al8Ag8/Fe65.4C7.1Si3.3B5.5P8.7Cr8Al2 (nanocrystalline)hot-pressing method3.4 ± 0.1[168]
Cu85Zr10Al5arc melter180021601.3[174]
W/Ni59Zr20Ti16Si2Sn3Milling/spark plasma sintering173024092.6[176]
Ti44Zr20Nb12Cu5Be19Arc melting/annealing14388.6[181]
MgxZrxCuxSputtering170052%[183]
Ti-40Nb (Ti74.4Nb25.6)Copper mold casting/melt spinning method544 ± 6628 ± 7[184]

6.2. Dendritic and Quasicrystal Metallic Glass Composite

G.H. Duan et al. showed that the shear bands can be terminated in two ways for dendritic phases. The first is the change in the position of the direction of the shear bands beside the boundaries of dendrites and glass matrix. The second one is the dendrites blocked the shear bands in the glass matrix [196]. Distinct microstructure in FeCrMoBC alloy with different cooling rates can be achieved by laser-directed energy deposition. The varying cooling rates generate the microstructure of soft crystalline dendrites with an amorphous matrix and lead to an intrinsic toughening mechanism [197]. J.I. Lee et al. prepared a metallic glass and alumina composite with the dendritic structure and showed that the composite at 800 °C showed the highest compressive strength of 2490 ± 65 MPa between the samples [198]. A new deformation method that pulls apart was discovered in in situ dendrites. In this method, dendrites are inserted and then torn apart in the toughened amorphous matrix during the plastic flow [199]. Dendritic ß-Ti phase and martensitic transformation increased the strain hardening of the composite. The author also studied the different volume fraction of ß-Ti dendrites and their impact on the composite and concluded that the combination of ß-Ti and 38% volume fraction demonstrates high ductility and strength [200]. The in situ β-Ti dendrite phase in Ti-MG stabilizes the plastic flow, and the generation of nucleation sites of the multiple shear bands leads to enhanced plasticity and strength (Table 8) of the composite [201]. S. Yang et al. concluded that the intermediate glassy phase at the nanoscale that is deformed homogenously could be related to the mechanical behavior of the Ti68Cu13.2Pd5.6Sn1.2Nb12 alloy [202]. The deformation-induced martensitic transformation in the B2–ZrCo dendrite in Zr–Co–Al (Z1) displayed high strength and plasticity. High strength and low plasticity were also found in Zr–Co–Al (Z5), due to the presence of a large eutectic inter-dendritic phase in the crystalline phases [203]. N. Hua et al. showed that high-strength ultrafine eutectic–dendrite composites with a high range of ductility and strength can be attained without rapid solidification and an injection mold casting procedure [204]. The Ti16.6Nb6Co5.1Cu6.5Al alloy has better potential than commercial Ti alloys in biomedical applications [205]. The mechanical properties in Ti–Zr–Nb–Cu–Be were increased and showed structural stability after annealing, due to slight β-thermal relaxation [206]. The addition of Nb in the Ti–Zr–Cu–Pd–Sn metallic glass and formation of the β-Ti phase increased the plasticity (Figure 10) and work- hardening feature [207]. The figure shows the increase in plasticity with the addition of Nb.
A 2% addition of the Si to Ti metallic glass produced a dendritic, which enhances the mechanical properties of the composite [208]. J. Cui et al. showed that the plasticity of the Ti50Zr20Nb12Cu5Be13 is enhanced by the emergence of the plastic dimple fracture and multiple shear bands [209]. The solid solution strengthening effect is caused by the impurities that are immersed in the ß titanium phase in the Ti48Zr20Nb12Cu5Be15. The author also concluded that the addition of a high amount of oxygen caused degradation in the composite [210]. During solid solution strengthening, the strength and hardness of the dendrites were increased in the Ta-based metallic glass, resulting in the generation of multiple shear bands, which improve the plasticity of the composite [211]. A new approach to form a high-strength Ti-based composite was proposed by inducing a brittle inter-dendritic phase (e.g., Co, Cr, or Cu) until the alloy exhibited β-Ti morphology completely [212]. H. S. Arora et al. showed that the friction stir processing procedure can be used for localized microstructural transformation and homogeneous distribution in the composite [213]. The strain rate of the Zr39.6Ti33.9Nb7.6Cu6.4Be12.5 composite is affected by the dispersion of the fracture and shear bands [214]. The withdrawal velocity in Zr-based metallic glass can affect the morphology and mechanical properties. The volume fraction in the crystalline phases is accountable for the improvement in the plasticity of the composite [215]. The formation of dendritic B2 CuZr improves the yield strength and ductility of the Zr metallic glass-based composite [216]. The shear band formation can be reduced by inducing quasicrystals in the amorphous matrix. They also hinder the formation of shear bands and cause the formation of multiple shear bands, increasing the plasticity [217]. The icosahedral quasistatic phase in the Mg matrix increases the plasticity (Table 9) and overall strength of the composite and has a chemical correlation between the quasicrystal phase and the metallic glass matrix [218]. The formation of the I-phase favored by oxygen impurity in the (Zr65Cu15Al10Ni10)90Nb10 leads to the reducing of the undercooling in the alloy melt, and during cooling, the long-range order of the icosahedral order is easily trapped [219].
Table 8. Mechanical properties of dendritic of metallic glass composites glass composite.
Table 8. Mechanical properties of dendritic of metallic glass composites glass composite.
Chemical CompositionPreparation MethodYield Strength (MPa)Ultimate Tensile StrengthPlasticityReference
Ti48Zr20Nb12Cu5Be15 metallic glasscold crucible levitation melting/copper mold casting1160131010.7%[199]
BT80copper mold casting1046 ± 171326 ± 159.1 ± 0.3[200]
(Ti51Zr38Nb11)71Cu9Be20suction casting method20538.06[201]
Ti68Cu13.2Pd5.6Sn1.2Nb12suction casting130023.7[202]
Zr54Co35Al11 (Z1)Suction cast/Quenching1323 ± 1028.7± 0.1[203]
Zr54Co35Al11 (Z5)Suction cast/Quenching1950 ± 103.0 ± 0.1[203]
Ni90Hf10arc-melting191034.4[204]
Ti16.6Nb6Co5.1Cu6.5AlArc melting/Casting1230 ± 5011 ± 1%[205]
Ti48Zr20Nb12Cu5Be15arc-melting/copper casting159934%[206]
Ti56Zr6Cu19.8Pd8.4Sn1.8Nb8Suction casting/copper molding1690268020[207]
(Ti0.425Cu0.425Ni0.075Zr0.075)1−xSixmelt-spinning/suction casting1497 ± 5010.3 ± 0.1[208]
Ti38.8Zr28.8Cu6.2Be16.2Nb10(Ta0.5 (VF))Melt Spinning/Copper-mold suction casting1517261020.3[209]
Ti68.8Nb13.6Cr5.1Co6Al6.5Melt spinning/Casting method1100 ± 201290 ± 5021 ± 3[212]
Zr60.0Ti14.7Nb5.3Cu5.6−Ni4.4Be10.0 (at%)Arc melting/the semisolid treatment plus Bridgman solidification1260205016.7[215]
cast Cu47.5Zr47.5Al5 (series 130)Arc melter/copper mold suction casting1311162813.94[216]
Table 9. Mechanical properties of quasicrystal of metallic glass composites glass composite.
Table 9. Mechanical properties of quasicrystal of metallic glass composites glass composite.
Chemical CompositionPreparation MethodYield Strength (MPa)Fracture Strength (MPa)Plastic StrainReference
Mg66Zn29Ca4YSemi-solid processing8508700.6%[218]

Summary

Metallic glass composites are generally prepared by either ex situ or in situ processes. In the ex situ process, secondary phases are produced by ball milling or melt infiltration. In situ processes are generally heat treatment for secondary phases. The formation of semi-crystalline, nanocrystalline, and quasi-crystalline particles can produce a metallic glass composite with high properties. Crystalline materials are ductile, and metallic glass lacks plasticity. The combination of both in a crystalline/metallic glass composite enhances the mechanical properties.

7. Conclusions and Prospects for the Future

Metallic glass is amorphous and is known for its mechanical properties, such as high toughness and strength, but lacks plasticity. It can be used in various applications [10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31] and is prepared by techniques such as melt-spinning, liquid squat quenching, powder metallurgy, magnetron sputtering, and pulsed laser quenching. The formation of metallic glass depends on the supercooled region and the critical cooling rate (Rc). Metallic glass has excellent mechanical properties as compared to crystalline materials but has low plasticity. This can be overcome by the introduction of secondary materials to form a metallic glass composite. The main aim of this review is to provide a brief overview; the background on metallic glass, its properties, and its application are discussed. Biocompatible composite and metallic glass based on Ti, Mg, Fe, and Sr are discussed. They have high strength and can be used in bioimplants and stents. The ex situ composites, fiber-reinforced metallic glass composites, are a metallic glass polymer and are also discussed. These composites have increased physical–mechanical properties. Biomaterial composites and metallic glass based on Ti, Mg, and Fe, are discussed. Metallic glass composite could be used as degradable bioimplants. Ti–Zr–Cu–Pd–Sn shows better workability than the commercial Ti and Ti–6Al–4V bioimplants. Biocompatibility and corrosion resistance can be increased by adding boron to Ti-based metallic glass. In general, Ti-based metallic glass and its composite show better cytocompatibility, corrosion resistance, and mechanical properties. Fiber-reinforced metallic glass composites and metallic glass/polymers composites are also discussed. Zr-based metallic glass is used in the musculoskeletal system. Zr–Ti–Cu–Ni–Be metallic glasses are used in ceramic films to increase their biocompatibility and tribological properties. Zr–Cu–Ni–Al composites have better plasticity and are used in load-bearing implants. The formation of the (Zr2Cu) nanocrystals leads to passivation, and an increase in the Y% increases cell activity. The Mg–Zn–Ca/PCL composite shows biocompatibility. PCL/nHA/Mg (MG) controls the Mg ions release. Mg–Zn–Ca ribbons have better resistance, biocompatibility, and tribological properties. It was also shown that the corrosion resistance of Mg–Zn–Ca is better than the commercial AZ31B and Mg. The addition of the Ag in Mg–Zn–Ca increases the corrosion and controls the hydrogen evolution. The cooling rate of the Mg–Zn–Ca metallic glass also influences the corrosion rate in the saline solution. Fe-based composites have potential in hyperthermia and tumor therapy. The formation of the carbon fiber/Ti–Zr–Be–Fe–Cu and W/(Zr–Ti–Cu–Ni) increases the mechanical properties, and the addition of Be in the Zr–Ti–Ni–Cu increases the adhesion of the composite. The addition of the Nb increases the interfacial strength between the W/MG composite. Recent research was conducted on MG and polymer composites. Some examples are Mg(MG)/PET, Mg(MG)/HDFE, Zr(MG)/PAN, and Cu(MG)/PTFE. They are used in structural and industrial applications. Various binders, such as silane, can be used to improve the interaction between metallic glass and polymer. The addition of the secondary reinforcement and introduction of the secondary phases increases the mechanical properties of the metallic glass composites. These composites are generally prepared by mechanical alloying or heat treatment procedures. The introduction of multiple shear bands and the formation of nano-crystalline, dendritic, and quasicrystal phases increase the plasticity of the metallic glass composite.
There is a great scope for composites based on metallic glass. Both physical and chemical properties can be increased by it. In the future, metallic glass composites have bright prospects in engineering and industrial applications.

Author Contributions

A.S., conceptualization, investigation, writing—original draft preparation; V.Z., review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

The reported study was funded by RFBR, project number 20-32-90041. V.Z. gratefully acknowledges the financial support made within the Ministry of Science and Higher Education of the Russian Federation in the framework of the Increase Competitiveness Program of NUST «MISiS» (grant number K3-2015-036).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Klement, W.; Willens, R.H.; Duwez, P. Non-crystalline Structure in Solidified Gold–Silicon Alloys. Nat. Cell Biol. 1960, 187, 869–870. [Google Scholar] [CrossRef]
  2. Ruhl, R.C. Cooling rates in splat cooling. Mater. Sci. Eng. 1967, 1, 313–320. [Google Scholar] [CrossRef]
  3. Chen, H.; Turnbull, D. Formation, stability and structure of palladium-silicon based alloy glasses. Acta Met. 1969, 17, 1021–1031. [Google Scholar] [CrossRef]
  4. Inoue, A.; Zhang, T.; Masumoto, T. Production of Amorphous Cylinder and Sheet of La55Al25Ni20 Alloy by a Metallic Mold Casting Method. Mater. Trans. JIM 1990, 31, 425–428. [Google Scholar] [CrossRef] [Green Version]
  5. Lee, M.C.; Johnson, W.L. A two-dimensional phase separation on the spherical surface of the metallic glass Au55Pb22.5Sb22.5. Appl. Phys. Lett. 1982, 41, 1054–1056. [Google Scholar] [CrossRef]
  6. Drehman, A.J.; Greer, A.L. Kinetics of crystal nucleation and growth in Pd40Ni40P20 glass. Acta Met. 1984, 32, 323–332. [Google Scholar] [CrossRef]
  7. Budhani, R.C.; Goel, T.C.; Chopra, K.L. Melt-spinning technique for preparation of metallic glasses. Bull. Mater. Sci. 1982, 4, 549–561. [Google Scholar] [CrossRef] [Green Version]
  8. Peker, A.; Johnson, W.L. A highly processable metallic glass: Zr41.2Ti13.8Cu12.5Ni10.0Be22.5. Appl. Phys. Lett. 1993, 63, 2342–2344. [Google Scholar] [CrossRef] [Green Version]
  9. Suryanarayana, C.; Inoue, A. Bulk Metallic Glasses, 2nd ed.; Routledge: London, UK, 2011; ISBN 9781420085976. [Google Scholar]
  10. Sohrabi, N.; Panikar, R.S.; Jhabvala, J.; Buch, A.R.; Mischler, S.; Logé, R.E. Laser coating of a Zr-based metallic glass on an aluminum substrate. Surf. Coat. Technol. 2020, 400, 126223. [Google Scholar] [CrossRef]
  11. Marattukalam, J.J.; Pacheco, V.; Karlsson, D.; Riekehr, L.; Lindwall, J.; Forsberg, F.; Jansson, U.; Sahlberg, M.; Hjörvarsson, B. Development of process parameters for selective laser melting of a Zr-based bulk metallic glass. Addit. Manuf. 2020, 33, 101124. [Google Scholar] [CrossRef]
  12. Hu, L.; Liu, X.; Chen, T.; Le, G.; Li, J.; Qu, F.; Zhou, Y.; Qi, L.; Wang, D. Characterization of laser cladded Zr–Cu–Ni–Al in-situ metallic glass matrix composite coatings with enhanced corrosion-resistance. Vacuum 2021, 185, 1–7. [Google Scholar] [CrossRef]
  13. Li, P.; Wang, J.J.; Cheng, J.L.; Li, F.; Zhao, W.; Chen, G. Investigation of the effects of Al on the glass forming ability of Zr-Cu-Ni-Al alloys through their solidification characteristics. Intermetallics 2019, 109, 105–109. [Google Scholar] [CrossRef]
  14. Hubert, B.; Yiu, P.; Hu, C.-C.; Chu, J.P. Fe-based thin film metallic glass as an activator of peroxymonosulfate for azo dye degradation. Surf. Coat. Technol. 2021, 412, 127031. [Google Scholar] [CrossRef]
  15. Sun, H.; Zheng, H.; Yang, X. Efficient degradation of orange II dye using Fe-based metallic glass powders prepared by commercial raw materials. Intermetallics 2021, 129, 107030. [Google Scholar] [CrossRef]
  16. Tsui, H.-P.; Lee, P.-H.; Yeh, C.-C.; Hung, J.-C. Ultrasonic vibration-assisted electrical discharge machining on Fe-based metallic glass by adding conductive powder. Procedia CIRP 2020, 95, 425–430. [Google Scholar] [CrossRef]
  17. Fan, J.T.; Zhang, Z.F.; Mao, S.X.; Shen, B.L.; Inoue, A. Deformation and fracture behaviors of Co-based metallic glass and its composite with dendrites. Intermetallics 2009, 17, 445–452. [Google Scholar] [CrossRef]
  18. Schnabel, V.; Köhler, M.; Evertz, S.; Gamcova, J.; Bednarcik, J.; Music, D.; Raabe, D.; Schneider, J.M. Revealing the relationships between chemistry, topology and stiffness of ultrastrong Co-based metallic glass thin films: A combinatorial approach. Acta Mater. 2016, 107, 213–219. [Google Scholar] [CrossRef]
  19. Wu, W.; Jiang, J.; Li, G.; Fuh, J.Y.H.; Jiang, H.; Gou, P.; Zhang, L.; Liu, W.; Zhao, J. Ultrasonic additive manufacturing of bulk Ni-based metallic glass. J. Non-Cryst. Solids 2019, 506, 1–5. [Google Scholar] [CrossRef]
  20. Bahrami, F.; Amini, R.; Taghvaei, A.H. Microstructure and corrosion behavior of electrodeposited Ni-based nanocomposite coatings reinforced with Ni 60 Cr 10 Ta 10 P 16 B 4 metallic glass particles. J. Alloys Compd. 2017, 714, 530–536. [Google Scholar] [CrossRef]
  21. Ma, X.; Ma, J.; Bian, X.; Tong, X.; Han, D.; Jia, Y.; Wu, S.; Zhang, N.; Geng, C.; Li, P.; et al. The role of nano-scale elastic heterogeneity in mechanical and tribological behaviors of a Cu–Zr based metallic glass thin films. Intermetallics 2021, 133, 107159. [Google Scholar] [CrossRef]
  22. Kim, S.-Y.; Park, G.-H.; Kim, H.-A.; Lee, A.-Y.; Oh, H.-R.; Lee, C.-W.; Lee, M.-H. Micro-deposition of Cu-based metallic glass wire by direct laser melting process. Mater. Lett. 2017, 202, 1–4. [Google Scholar] [CrossRef]
  23. Yang, C.; Lu, L.-M.; Zhao, Z.-W.; Li, J.-H.; Gong, F.; Ma, J. Micro thermoplastic forming of a Pd-based metallic glass: Theory and applications. J. Iron Steel Res. Int. 2017, 24, 378–384. [Google Scholar] [CrossRef]
  24. Pineda, E.; Serrano, J.; Bruna, P.; Crespo, D. Fragility measurement of Pd-based metallic glass by dynamic mechanical analysis. J. Alloys Compd. 2010, 504, S215–S218. [Google Scholar] [CrossRef]
  25. Wang, Y.B.; Xie, X.H.; Li, H.F.; Wang, X.L.; Zhao, M.Z.; Zhang, E.W.; Bai, Y.J.; Zheng, Y.F.; Qin, L. Biodegradable CaMgZn bulk metallic glass for potential skeletal application. Acta Biomater. 2011, 7, 3196–3208. [Google Scholar] [CrossRef]
  26. Khan, M.M.; Nemati, A.; Rahman, Z.U.; Shah, U.H.; Asgar, H.; Haider, W. Recent Advancements in Bulk Metallic Glasses and Their Applications: A Review. Crit. Rev. Solid State Mater. Sci. 2018, 43, 233–268. [Google Scholar] [CrossRef]
  27. Kumar, G.P.; Tavakoli, R.; Cui, F.; Jafary-Zadeh, M. Feasibility of using bulk metallic glass for self-expandable stent applications. J. Biomed. Mater. Res. Part B Appl. Biomater. 2016, 105, 1874–1882. [Google Scholar] [CrossRef]
  28. Phan, T.A.; Hara, M.; Oguchi, H.; Kuwano, H. Current sensors using Fe–B–Nd–Nb magnetic metallic glass micro-cantilevers. Microelectron. Eng. 2015, 135, 28–31. [Google Scholar] [CrossRef]
  29. Lin, Y.-C.; Tsai, Y.-C.; Ono, T.; Liu, P.; Esashi, M.; Gessner, T.; Chen, M. Metallic Glass as a Mechanical Material for Microscanners. Adv. Funct. Mater. 2015, 25, 5677–5682. [Google Scholar] [CrossRef]
  30. Liu, W.; Zhang, H.; Shi, J.-A.; Wang, Z.; Song, C.; Wang, X.; Lu, S.; Zhou, X.; Gu, L.; Louzguine-Luzgin, D.; et al. A room-temperature magnetic semiconductor from a ferromagnetic metallic glass. Nat. Commun. 2016, 7, 13497. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Hasannaeimi, V.; Mukherjee, S. Noble-Metal based Metallic Glasses as Highly Catalytic Materials for Hydrogen Oxidation Reaction in Fuel Cells. Sci. Rep. 2019, 9, 1–8. [Google Scholar] [CrossRef] [Green Version]
  32. Monfared, A.; Vali, H.; Faghihi, S. Biocorrosion and biocompatibility of Zr-Cu-Fe-Al bulk metallic glasses. Surf. Interface Anal. 2013, 45, 1714–1720. [Google Scholar] [CrossRef]
  33. Filipecka-Szymczyk, K.; Pawlik, P.; Filipecki, J. The effect of annealing on magnetic properties, phase structure and evolution of free volumes in Pr-Fe-B-W metallic glasses. J. Alloys Compd. 2017, 694, 228–234. [Google Scholar] [CrossRef]
  34. Wang, Z.; Georgarakis, K.; Nakayama, K.S.; Li, Y.; Tsarkov, A.A.; Xie, G.; Dudina, D.; Louzguine-Luzgin, D.V.; Yavari, A.R. Microstructure and mechanical behavior of metallic glass fiber-reinforced Al alloy matrix composites. Sci. Rep. 2016, 6, 1–11. [Google Scholar] [CrossRef] [Green Version]
  35. Davies, H.A.; Aucote, J.; Hull, J.B. Amorphous Nickel produced by Splat Quenching. Nat. Phys. Sci. 1973, 246, 13–14. [Google Scholar] [CrossRef]
  36. Ding, S.; Liu, Y.; Li, Y.; Liu, Z.; Sohn, S.; Walker, F.J.; Schroers, J. Combinatorial development of bulk metallic glasses. Nat. Mater. 2014, 13, 494–500. [Google Scholar] [CrossRef]
  37. Lin, C.-J. Fe-B glasses formed by picosecond pulsed laser quenching. Appl. Phys. Lett. 1982, 41, 721. [Google Scholar] [CrossRef]
  38. Fan, G.J.; Fu, L.F.; Qiao, D.C.; Choo, H.; Liaw, P.K.; Browning, N.D.; Löffler, J.F. Effect of microalloying on the glass-forming ability of Cu60Zr30Ti10 bulk metallic glass. J. Non-Cryst. Solids 2007, 353, 4218–4222. [Google Scholar] [CrossRef]
  39. Mukherjee, S.; Schroers, J.; Zhou, Z.; Johnson, W.L.; Rhim, W.-K. Viscosity and specific volume of bulk metallic glass-forming alloys and their correlation with glass forming ability. Acta Mater. 2004, 52, 3689–3695. [Google Scholar] [CrossRef]
  40. Park, E.S.; Na, J.H.; Kim, D.H. Correlation between fragility and glass-forming ability/plasticity in metallic glass-forming alloys. Appl. Phys. Lett. 2007, 91, 1–4. [Google Scholar] [CrossRef]
  41. Bordeenithikasem, P.; Liu, J.; Kube, S.A.; Li, Y.; Ma, T.; Scanley, B.E.; Broadbridge, C.C.; Vlassak, J.J.; Singer, J.P.; Schroers, J. Determination of critical cooling rates in metallic glass forming alloy libraries through laser spike annealing. Sci. Rep. 2017, 7, 1–9. [Google Scholar] [CrossRef] [Green Version]
  42. Hildal, K.; Sekido, N.; Perepezko, J.H. Critical cooling rate for Fe48Cr15Mo14Y2C15B6 bulk metallic glass formation. Intermetallics 2006, 14, 898–902. [Google Scholar] [CrossRef]
  43. Takeuchi, A.; Inoue, A. Quantitative evaluation of critical cooling rate for metallic glasses. Mater. Sci. Eng. A 2001, 304-306, 446–451. [Google Scholar] [CrossRef]
  44. Busch, R.; Kim, Y.J.; Johnson, W.L. Thermodynamics and kinetics of the undercooled liquid and the glass transition of the Zr41.2Ti13.8Cu12.5Ni10.0Be22.5 alloy. J. Appl. Phys. 1995, 77, 4039–4043. [Google Scholar] [CrossRef] [Green Version]
  45. Liu, N.; Ma, T.; Liao, C.; Liu, G.; Mota, R.M.O.; Liu, J.; Sohn, S.; Kube, S.; Zhao, S.; Singer, J.P.; et al. Combinatorial measurement of critical cooling rates in aluminum-base metallic glass forming alloys. Sci. Rep. 2021, 11, 3903. [Google Scholar] [CrossRef]
  46. Yodoshi, N.; Yamada, R.; Kawasaki, A.; Makino, A. Evaluation of critical cooling rate of Fe76Si9B10P5 metallic glass by containerless solidification process. J. Alloys Compd. 2015, 643, S2–S7. [Google Scholar] [CrossRef]
  47. Nishiyama, N.; Inoue, A. Supercooling investigation and critical cooling rate for glass formation in Pd–Cu–Ni–P alloy. Acta Mater. 1999, 47, 1487–1495. [Google Scholar] [CrossRef]
  48. Pogatscher, S.; Uggowitzer, P.J.; Löffler, J.F. In-situ probing of metallic glass formation and crystallization upon heating and cooling via fast differential scanning calorimetry. Appl. Phys. Lett. 2014, 104, 251908. [Google Scholar] [CrossRef] [Green Version]
  49. Shen, Y.; Li, Y.; Tsai, H.-L. Evolution of crystalline phase during laser processing of Zr-based metallic glass. J. Non-Cryst. Solids 2018, 481, 299–305. [Google Scholar] [CrossRef]
  50. Castellero, A.; Fiore, G.; Van Steenberge, N.; Battezzati, L. Processing a Fe67Mo4.5Cr2.3Al2Si3C7P8.7B5.5 metallic glass: Experimental and computed TTT and CCT curves. J. Alloys Compd. 2020, 843, 156061. [Google Scholar] [CrossRef]
  51. Liao, J.P.; Yang, B.J.; Zhang, Y.; Lu, W.Y.; Gu, X.J.; Wang, J.Q. Evaluation of glass formation and critical casting diameter in Al-based metallic glasses. Mater. Des. 2015, 88, 222–226. [Google Scholar] [CrossRef]
  52. Ashby, M.F.; Greer, A.L. Metallic glasses as structural materials. Scr. Mater. 2006, 54, 321–326. [Google Scholar] [CrossRef]
  53. Greer, A.L. Metallic glasses…on the threshold. Mater. Today 2009, 12, 14–22. [Google Scholar] [CrossRef]
  54. Fornell, J.; Van Steenberge, N.; Varea, A.; Rossinyol, E.; Pellicer, E.; Suriñach, S.; Baró, M.D.; Sort, J. Enhanced mechanical properties and in vitro corrosion behavior of amorphous and devitrified Ti40Zr10Cu38Pd12 metallic glass. J. Mech. Behav. Biomed. Mater. 2011, 4, 1709–1717. [Google Scholar] [CrossRef] [PubMed]
  55. Peter, W.H.; Buchanan, R.A.; Liu, C.T.; Liaw, P.K.; Morrison, M.L.; Horton, J.A.; Carmichael, C.A.; Wright, J.L. Localized corrosion behavior of a zirconium-based bulk metallic glass relative to its crystalline state. Intermetallics 2002, 10, 1157–1162. [Google Scholar] [CrossRef]
  56. Madge, S.V.; Caron, A.; Gralla, R.; Wilde, G.; Mishra, S.K. Novel W-based metallic glass with high hardness and wear resistance. Intermetallics 2014, 47, 6–10. [Google Scholar] [CrossRef]
  57. Wang, W.H. Bulk Metallic Glasses with Functional Physical Properties. Adv. Mater. 2009, 21, 4524–4544. [Google Scholar] [CrossRef]
  58. Sharma, P.; Kaushik, N.; Kimura, H.; Saotome, Y.; Inoue, A. Nano-fabrication with metallic glass—An exotic material for nano-electromechanical systems. Nanotechnology 2007, 18, 035302. [Google Scholar] [CrossRef]
  59. Nieh, T.G.; Wadsworth, J.; Liu, C.T.; Ohkubo, T.; Hirotsu, Y. Plasticity and structural instability in a bulk metallic glass deformed in the supercooled liquid region. Acta Mater. 2001, 49, 2887–2896. [Google Scholar] [CrossRef]
  60. Qiao, J.W.; Zhang, Y.A.; Jia, H.L.; Yang, H.J.; Liaw, P.K.; Xu, B.S. Tensile softening of metallic-glass-matrix composites in the supercooled liquid region. Appl. Phys. Lett. 2012, 100, 9–13. [Google Scholar] [CrossRef]
  61. Zhang, Y.; Greer, A.L. Correlations for predicting plasticity or brittleness of metallic glasses. J. Alloys Compd. 2007, 434-435, 2–5. [Google Scholar] [CrossRef]
  62. Khonik, V.A. Understanding of the Structural Relaxation of Metallic Glasses within the Framework of the Interstitialcy Theory. Metals 2015, 5, 504–529. [Google Scholar] [CrossRef] [Green Version]
  63. Schuh, C.A.; Hufnagel, T.C.; Ramamurty, U. Mechanical behavior of amorphous alloys. Acta Mater. 2007, 55, 4067–4109. [Google Scholar] [CrossRef]
  64. Dyre, J.C. Colloquium: The glass transition and elastic models of glass-forming liquids. Rev. Mod. Phys. 2006, 78, 953–972. [Google Scholar] [CrossRef] [Green Version]
  65. Yu, H.-B.; Shen, X.; Wang, Z.; Gu, L.; Wang, W.H.; Bai, H.Y. Tensile Plasticity in Metallic Glasses with PronouncedβRelaxations. Phys. Rev. Lett. 2012, 108, 1–5. [Google Scholar] [CrossRef] [PubMed]
  66. Langer, J.S. Shear-transformation-zone theory of plastic deformation near the glass transition. Phys. Rev. E 2008, 77, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Greer, A.; Cheng, Y.Q.; Ma, E. Shear bands in metallic glasses. Mater. Sci. Eng. R: Rep. 2013, 74, 71–132. [Google Scholar] [CrossRef]
  68. Maaß, R.; Klaumünzer, D.; Preiß, E.I.; Derlet, P.M.; Löffler, J.F. Single shear-band plasticity in a bulk metallic glass at cryogenic temperatures. Scr. Mater. 2012, 66, 231–234. [Google Scholar] [CrossRef]
  69. Meduri, C.; Hasan, M.; Adam, S.; Kumar, G. Effect of temperature on shear bands and bending plasticity of metallic glasses. J. Alloys Compd. 2018, 732, 922–927. [Google Scholar] [CrossRef]
  70. Shuai, C.; Li, S.; Peng, S.; Feng, P.; Lai, Y.; Gao, C. Biodegradable metallic bone implants. Mater. Chem. Front. 2019, 3, 544–562. [Google Scholar] [CrossRef]
  71. Li, H.F.; Zheng, Y.F. Recent advances in bulk metallic glasses for biomedical applications. Acta Biomater. 2016, 36, 1–20. [Google Scholar] [CrossRef] [PubMed]
  72. Tsai, P.H.; Lin, Y.Z.; Li, J.B.; Jian, S.R.; Jang, J.S.C.; Li, C.; Chu, J.P.; Huang, J.C.C. Sharpness improvement of surgical blade by means of ZrCuAlAgSi metallic glass and metallic glass thin film coating. Intermetallics 2012, 31, 127–131. [Google Scholar] [CrossRef]
  73. Huang, L.; Pu, C.; Fisher, R.K.; Mountain, D.J.H.; Gao, Y.; Liaw, P.K.; Zhang, W.; He, W. A Zr-based bulk metallic glass for future stent applications: Materials properties, finite element modeling, and in vitro human vascular cell response. Acta Biomater. 2015, 25, 356–368. [Google Scholar] [CrossRef] [Green Version]
  74. Aliyu, A.A.A.; Abdul-Rani, A.M.; Ginta, T.L.; Prakash, C.; Axinte, E.; Fua-Nizan, R. Investigation of nanoporosities fabricated on metallic glass surface by hydroxyapatite mixed EDM for orthopedic application. Malays. J. Fundam. Appl. Sci. 2017, 13, 523–528. [Google Scholar] [CrossRef]
  75. Wu, Z.; Du, P.; Xiang, T.; Li, K.; Xie, G. Ti-based bulk metallic glass implantable biomaterial with adjustable porosity produced by a novel pressure regulation method in spark plasma sintering. Intermetallics 2021, 131, 107105. [Google Scholar] [CrossRef]
  76. Xie, G.; Kanetaka, H.; Kato, H.; Qin, F.; Wang, W. Porous Ti-based bulk metallic glass with excellent mechanical properties and good biocompatibility. Intermetallics 2019, 105, 153–162. [Google Scholar] [CrossRef]
  77. Wu, H.; Liang, L.; Lan, X.; Yin, Y.; Song, M.; Li, R.; Liu, Y.; Yang, H.; Liu, L.; Cai, A.; et al. Tribological and biological behaviors of laser cladded Ti-based metallic glass composite coatings. Appl. Surf. Sci. 2020, 507, 145104. [Google Scholar] [CrossRef]
  78. Rajan, S.T.; Das, M.; Kumar, P.S.; Arockiarajan, A.; Subramanian, B. Biological performance of metal metalloid (TiCuZrPd:B) TFMG fabricated by pulsed laser deposition. Colloids Surf. B: Biointerfaces 2021, 202, 111684. [Google Scholar] [CrossRef] [PubMed]
  79. Lin, C.H.; Chen, C.-H.; Huang, Y.S.; Huang, C.H.; Huang, J.C.; Jang, J.S.C.; Lin, Y.S. In vivo investigations and cytotoxicity tests on Ti/Zr-based metallic glasses with various Cu contents. Mater. Sci. Eng. C 2017, 77, 308–317. [Google Scholar] [CrossRef]
  80. Liu, Y.; Wang, H.-J.; Pang, S.-J.; Zhang, T. Ti–Zr–Cu–Fe–Sn–Si–Ag–Ta bulk metallic glasses with good corrosion resistance as potential biomaterials. Rare Met. 2020, 39, 688–694. [Google Scholar] [CrossRef]
  81. Li, P.; Meng, L.; Wang, S.; Wang, K.; Sui, Q.; Liu, L.; Zhang, Y.; Yin, X.; Zhang, Q.; Wang, L. In Situ Formation of Ti47Cu38Zr7.5Fe2.5Sn2Si1Nb2 Amorphous Coating by Laser Surface Remelting. Materials 2019, 12, 3660. [Google Scholar] [CrossRef] [Green Version]
  82. Yan, H.-M.; Liu, Y.; Pang, S.-J.; Zhang, T. Glass formation and properties of Ti-based bulk metallic glasses as potential biomaterials with Nb additions. Rare Met. 2018, 37, 831–837. [Google Scholar] [CrossRef]
  83. Liens, A.; Etiemble, A.; Rivory, P.; Balvay, S.; Pelletier, J.-M.; Cardinal, S.; Fabrègue, D.; Kato, H.; Steyer, P.; Munhoz, T.; et al. On the Potential of Bulk Metallic Glasses for Dental Implantology: Case Study on Ti40Zr10Cu36Pd14. Materials 2018, 11, 249. [Google Scholar] [CrossRef] [Green Version]
  84. Abdi, S.; Oswald, S.; Gostin, P.-F.; Helth, A.; Sort, J.; Baró, M.D.; Calin, M.; Schultz, L.; Eckert, J.; Gebert, A. Designing new biocompatible glass-forming Ti75−xZr10NbxSi15 (x = 0, 15) alloys: Corrosion, passivity, and apatite formation. J. Biomed. Mater. Res. Part B Appl. Biomater. 2015, 104, 27–38. [Google Scholar] [CrossRef] [PubMed]
  85. Wang, Y.B.; Li, H.F.; Cheng, Y.F.; Zheng, Y.; Ruan, L.Q. In vitro and in vivo studies on Ti-based bulk metallic glass as potential dental implant material. Mater. Sci. Eng. C 2013, 33, 3489–3497. [Google Scholar] [CrossRef] [PubMed]
  86. Chen, N.; Shi, X.; Witte, R.; Nakayama, K.S.; Ohmura, K.; Wu, H.; Takeuchi, A.; Hahn, H.; Esashi, M.; Gleiter, H.; et al. A novel Ti-based nanoglass composite with submicron–nanometer-sized hierarchical structures to modulate osteoblast behaviors. J. Mater. Chem. B 2013, 1, 2568–2574. [Google Scholar] [CrossRef] [PubMed]
  87. Meagher, P.; O’Cearbhaill, E.D.; Byrne, J.H.; Browne, D.J. Bulk Metallic Glasses for Implantable Medical Devices and Surgical Tools. Adv. Mater. 2016, 28, 5755–5762. [Google Scholar] [CrossRef] [Green Version]
  88. Sawyer, V.; Tao, X.; Dong, H.; Dashtbozorg, B.; Li, X.; Sammons, R.; Dong, H.-S. Improving the Tribological Properties and Biocompatibility of Zr-Based Bulk Metallic Glass for Potential Biomedical Applications. Materials 2020, 13, 1960. [Google Scholar] [CrossRef]
  89. Abdul-Rani, A.M.; Aliyu, A.A.A.; Hastuty, S.; Ginta, T.L.; Rao, T.V.V.L.N.; Ali, S. Enhancing surface quality of Zr-Cu-Ni-Ti-Be through hydroxyapatite mixed EDM for potential orthopedic application. AIP Conf. Proc. 2018, 2035, 080010. [Google Scholar] [CrossRef]
  90. Hua, N.; Chen, W.; Wang, W.; Lu, H.; Ye, X.; Li, G.; Lin, C.; Huang, X. Tribological behavior of a Ni-free Zr-based bulk metallic glass with potential for biomedical applications. Mater. Sci. Eng. C 2016, 66, 268–277. [Google Scholar] [CrossRef]
  91. Subramanian, B.; Maruthamuthu, S.; Rajan, S.T. Biocompatibility evaluation of sputtered zirconium-based thin film metallic glass-coated steels. Int. J. Nanomed. 2015, 10, 17–29. [Google Scholar] [CrossRef] [Green Version]
  92. Huang, L.; Zhu, C.; Muntele, C.I.; Zhang, T.; Liaw, P.K.; He, W. Surface engineering of a Zr-based bulk metallic glass with low energy Ar- or Ca-ion implantation. Mater. Sci. Eng. C 2015, 47, 248–255. [Google Scholar] [CrossRef] [Green Version]
  93. Hua, N.; Huang, L.; Chen, W.; He, W.; Zhang, T. Biocompatible Ni-free Zr-based bulk metallic glasses with high-Zr-content: Compositional optimization for potential biomedical applications. Mater. Sci. Eng. C 2014, 44, 400–410. [Google Scholar] [CrossRef]
  94. Espallargas, N.; Aune, R.E.; Torres, C.; Papageorgiou, N.; Muñoz, A.I. Bulk metallic glasses (BMG) for biomedical applications—A tribocorrosion investigation of Zr55Cu30Ni5Al10 in simulated body fluid. Wear 2013, 301, 271–279. [Google Scholar] [CrossRef]
  95. Hua, N.; Huang, L.; He, W.; Pang, S.; Zhang, T. A Ni-free high-zirconium-based bulk metallic glass with enhanced plasticity and biocompatibility. J. Non-Cryst. Solids 2013, 376, 133–138. [Google Scholar] [CrossRef]
  96. Huang, L.; Cao, Z.; Meyer, H.M.; Liaw, P.K.; Garlea, E.; Dunlap, J.R.; Zhang, T.; He, W. Responses of bone-forming cells on pre-immersed Zr-based bulk metallic glasses: Effects of composition and roughness. Acta Biomater. 2011, 7, 395–405. [Google Scholar] [CrossRef]
  97. Sharma, A.; Kopylov, A.; Zadorozhnyy, M.; Stepashkin, A.; Kudelkina, V.; Wang, J.-Q.; Ketov, S.; Churyukanova, M.; Louzguine-Luzgin, D.; Sarac, B.; et al. Mg-Based Metallic Glass-Polymer Composites: Investigation of Structure, Thermal Properties, and Biocompatibility. Metals 2020, 10, 867. [Google Scholar] [CrossRef]
  98. Abisegapriyan, K.S.; Rajeshwari, A.; Kundu, S.; Subramanian, B. Magnesium glassy alloy laminated nanofibrous polymer as biodegradable scaffolds. J. Non-Cryst. Solids 2018, 502, 210–217. [Google Scholar] [CrossRef]
  99. Monfared, A.; Ghaee, A.; Ebrahimibarough, S. Preparation and characterization of crystallized and relaxed amorphous Mg-Zn-Ca alloy ribbons for nerve regeneration application. J. Non-Cryst. Solids 2018, 489, 71–76. [Google Scholar] [CrossRef]
  100. Hua, N.; Chen, W.; Wang, Q.; Guo, Q.; Huang, Y.; Zhang, T. Tribocorrosion behaviors of a biodegradable Mg65Zn30Ca5 bulk metallic glass for potential biomedical implant applications. J. Alloys Compd. 2018, 745, 111–120. [Google Scholar] [CrossRef]
  101. Baulin, O.; Fabrègue, D.; Kato, H.; Liens, A.; Wada, T.; Pelletier, J.-M. A new, toxic element-free Mg-based metallic glass for biomedical applications. J. Non-Cryst. Solids 2018, 481, 397–402. [Google Scholar] [CrossRef]
  102. Li, H.; He, W.; Pang, S.; Liaw, P.K.; Zhang, T. In vitro responses of bone-forming MC3T3-E1 pre-osteoblasts to biodegradable Mg-based bulk metallic glasses. Mater. Sci. Eng. C 2016, 68, 632–641. [Google Scholar] [CrossRef] [Green Version]
  103. Li, H.; Pang, S.; Liu, Y.; Liaw, P.K.; Zhang, T. In vitro investigation of Mg–Zn–Ca–Ag bulk metallic glasses for biomedical applications. J. Non-Cryst. Solids 2015, 427, 134–138. [Google Scholar] [CrossRef]
  104. Li, H.; Pang, S.; Liu, Y.; Sun, L.; Liaw, P.K.; Zhang, T. Biodegradable Mg–Zn–Ca–Sr bulk metallic glasses with enhanced corrosion performance for biomedical applications. Mater. Des. 2015, 67, 9–19. [Google Scholar] [CrossRef]
  105. Yu, H.-J.; Wang, J.-Q.; Shi, X.-T.; Louzguine-Luzgin, D.V.; Wu, H.-K.; Perepezko, J.H. Ductile Biodegradable Mg-Based Metallic Glasses with Excellent Biocompatibility. Adv. Funct. Mater. 2013, 23, 4793–4800. [Google Scholar] [CrossRef]
  106. Wang, J.; Huang, S.; Li, Y.; Wei, Y.; Xi, X.; Cai, K. Microstructure, mechanical and bio-corrosion properties of Mn-doped Mg–Zn–Ca bulk metallic glass composites. Mater. Sci. Eng. C 2013, 33, 3832–3838. [Google Scholar] [CrossRef]
  107. Esmaeili, A.; Ghaffari, S.A.; Nikkhah, M.; Ghaini, F.M.; Farzan, F.; Mohammadi, S. Biocompatibility assessments of 316L stainless steel substrates coated by Fe-based bulk metallic glass through electro-spark deposition method. Colloids Surf. B: Biointerfaces 2021, 198, 111469. [Google Scholar] [CrossRef]
  108. Li, S.; Wei, Q.; Li, Q.; Jiang, B.; Chen, Y.; Sun, Y. Development of Fe-based bulk metallic glasses as potential biomaterials. Mater. Sci. Eng. C 2015, 52, 235–241. [Google Scholar] [CrossRef]
  109. Yang, C.; Bian, X.; Qin, J.; Guo, T.; Zhao, S. Fabrication and hyperthermia effect of magnetic functional fluids based on amorphous particles. Appl. Surf. Sci. 2015, 330, 216–220. [Google Scholar] [CrossRef]
  110. Zohdi, H.; Shahverdi, H.R.; Hadavi, S.M.M. Effect of Nb addition on corrosion behavior of Fe-based metallic glasses in Ringer’s solution for biomedical applications. Electrochem. Commun. 2011, 13, 840–843. [Google Scholar] [CrossRef]
  111. Zhao, K.; Jiao, W.; Ma, J.; Gao, X.Q.; Wang, W.H. Formation and properties of strontium-based bulk metallic glasses with ultralow glass transition temperature. J. Mater. Res. 2012, 27, 2593–2600. [Google Scholar] [CrossRef]
  112. Lee, M.; Choi, Y.; Sugio, K.; Matsugi, K.; Sasaki, G. Effect of aluminum carbide on thermal conductivity of the unidirectional CF/Al composites fabricated by low pressure infiltration process. Compos. Sci. Technol. 2014, 97, 1–5. [Google Scholar] [CrossRef]
  113. Yin, H.L.; Liu, S.Q.; Zhao, L.C.; Cui, C.X.; Wang, X. Vacuum infiltration molding and mechanical property of short carbon fiber reinforced Ti-based metallic glass matrix composite. J. Mater. Process. Technol. 2021, 295, 117151. [Google Scholar] [CrossRef]
  114. Zhang, Z.; Kong, J.; Liu, X.; Song, X.; Dong, K. Preparation, microstructure and mechanical properties of tungsten fiber reinforced LaAlCuNi metallic glass matrix composites. Intermetallics 2021, 132, 107139. [Google Scholar] [CrossRef]
  115. Chen, S.; Li, W.Q.; Zhang, L.; Fu, H.M.; Li, Z.K.; Zhu, Z.W.; Li, H.; Zhang, H.W.; Wang, A.M.; Wang, Y.D. Dynamic compressive mechanical properties of the spiral tungsten wire reinforced Zr-based bulk metallic glass composites. Compos. Part B Eng. 2020, 199, 108219. [Google Scholar] [CrossRef]
  116. Zhou, J.; Wu, Y.; Wang, H.; Liu, X.; Jiang, S.; Wang, X.; Lu, Z. Work-hardenable Zr-based bulk metallic glass composites reinforced with ex-situ TiNi fibers. J. Alloys Compd. 2019, 806, 1497–1508. [Google Scholar] [CrossRef]
  117. Zhang, B.; Fu, H.; Zhang, H.; Mu, J.; Wang, M. Synthesis and property of short tungsten fibre/Zr-based metallic glass composite. Mater. Sci. Technol. 2019, 35, 1347–1354. [Google Scholar] [CrossRef]
  118. Chen, S.; Zhang, L.; Fu, H.M.; Li, Z.K.; Zhu, Z.W.; Li, H.; Zhang, H.W.; Wang, A.M.; Wang, Y.D.; Zhang, H.F. Compressive mechanical properties and failure modes of Zr-based bulk metallic glass composites containing tungsten springs. Mater. Des. 2018, 160, 652–660. [Google Scholar] [CrossRef]
  119. Hamill, L.; Nutt, S. Adhesion of metallic glass and epoxy in composite-metal bonding. Compos. Part B: Eng. 2018, 134, 186–192. [Google Scholar] [CrossRef]
  120. Chen, G.; Hao, Y.; Chen, X.; Hao, H. Compressive behaviour of tungsten fibre reinforced Zr-based metallic glass at different strain rates and temperatures. Int. J. Impact Eng. 2017, 106, 110–119. [Google Scholar] [CrossRef]
  121. Shamlaye, K.F.; Laws, K.J.; Ferry, M. Supercooled liquid fusion of carbon fibre-bulk metallic glass composites with superplastic forming properties. Scr. Mater. 2016, 111, 127–130. [Google Scholar] [CrossRef]
  122. Wang, Z.; Scudino, S.; Stoica, M.; Zhang, W.; Eckert, J. Al-based matrix composites reinforced with short Fe-based metallic glassy fiber. J. Alloys Compd. 2015, 651, 170–175. [Google Scholar] [CrossRef]
  123. Li, Z.K.; Fu, H.M.; Sha, P.F.; Zhu, Z.W.; Wang, A.M.; Li, H.; Zhang, H.W.; Zhang, H.F.; Hu, Z.Q. Atomic interaction mechanism for designing the interface of W/Zr-based bulk metallic glass composites. Sci. Rep. 2015, 5, 8967. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Chen, J.H.; Chen, Y.; Jiang, M.Q.; Chen, X.W.; Zhang, H.F.; Dai, L.H. On the compressive failure of tungsten fiber reinforced Zr-based bulk metallic glass composite. Int. J. Solids Struct. 2015, 69-70, 428–441. [Google Scholar] [CrossRef]
  125. Chen, J.H.; Chen, Y.; Jiang, M.Q.; Chen, X.W.; Zhang, H.F.; Dai, L.H. Dynamic shear punch behavior of tungsten fiber reinforced Zr-based bulk metallic glass matrix composites. Int. J. Impact Eng. 2015, 79, 22–31. [Google Scholar] [CrossRef] [Green Version]
  126. Zhang, B.; Fu, H.; Li, Z.; Zhu, Z.; Zhang, H.; Hu, Z. Anisotropic tensile properties of tungsten fiber reinforced Zr based metallic glass composites. Mater. Sci. Eng. A 2014, 619, 165–170. [Google Scholar] [CrossRef]
  127. Chen, J.H.; Chen, Y.; Jiang, M.Q.; Chen, X.W.; Fu, H.M.; Zhang, H.F.; Dai, L.H. Direct Observation on the Evolution of Shear Banding and Buckling in Tungsten Fiber Reinforced Zr-Based Bulk Metallic Glass Composite. Met. Mater. Trans. A 2014, 45, 5397–5408. [Google Scholar] [CrossRef] [Green Version]
  128. Khademian, N.; Gholamipour, R. Study on microstructure and fracture behavior of tungsten wire reinforced Cu-based and Zr-based bulk metallic glass matrix composites. J. Non-Cryst. Solids 2013, 365, 75–84. [Google Scholar] [CrossRef]
  129. Zhang, B.Y.; Chen, X.H.; Wang, S.S.; Lin, D.Y.; Hui, X.D. High strength tungsten wire reinforced Zr-based bulk metallic glass matrix composites prepared by continuous infiltration process. Mater. Lett. 2013, 93, 210–214. [Google Scholar] [CrossRef]
  130. Wang, Z.; Zhang, J.; Li, Y. Quasi-Static and Dynamic Compressive Characteristics of Tungsten Fiber/Zr-Based Bulk Metallic Glass Matrix Composites. J. Comput. Theor. Nanosci. 2012, 9, 1427–1430. [Google Scholar] [CrossRef]
  131. Rong, G.; Huang, D.W.; Yang, M.C. Penetrating behaviors of Zr-based metallic glass composite rods reinforced by tungsten fibers. Theor. Appl. Fract. Mech. 2012, 58, 21–27. [Google Scholar] [CrossRef]
  132. Jiang, F.; Chen, G.; Wang, Z.; Cao, Y.; Cheng, J.; Chen, G. Mechanical Properties of Tungsten Fiber Reinforced (Zr41.2Ti13.8Cu12.5Ni10Be22.5)100-xNbx Bulk Metallic Glass Composites. Rare Met. Mater. Eng. 2011, 40, 206–208. [Google Scholar] [CrossRef]
  133. Deng, S.T.; Diao, H.; Chen, Y.L.; Yan, C.; Zhang, H.F.; Wang, A.M.; Hu, Z.Q. Metallic glass fiber-reinforced Zr-based bulk metallic glass. Scr. Mater. 2011, 64, 85–88. [Google Scholar] [CrossRef]
  134. Hyeon, D.Y.; Nam, C.; Ham, S.S.; Hwang, G.T.; Yi, S.; Kim, K.T.; Park, K. Enhanced Energy Conversion Performance of a Magneto–Mechano–Electric Generator Using a Laminate Composite Made of Piezoelectric Polymer and Metallic Glass. Adv. Electron. Mater. 2021, 7, 1–9. [Google Scholar] [CrossRef]
  135. Zadorozhnyy, V.; Churyukanova, M.; Stepashkin, A.; Zadorozhnyy, M.; Sharma, A.; Moskovskikh, D.; Wang, J.; Shabanova, E.; Ketov, S.; Louzguine-Luzgin, D.; et al. Structure and Thermal Properties of an Al-Based Metallic Glass-Polymer Composite. Metals 2018, 8, 1037. [Google Scholar] [CrossRef] [Green Version]
  136. Zadorozhnyy, V.Y.; Gorshenkov, M.V.; Churyukanova, M.N.; Zadorozhnyy, M.Y.; Stepashkin, A.A.; Moskovskikh, D.O.; Ketov, S.V.; Zinnurova, L.K.; Sharma, A.; Louzguine-Luzgin, D.V.; et al. Investigation of structure and thermal properties in composite materials based on metallic glasses with small addition of polytetrafluoroethylene. J. Alloys Compd. 2017, 707, 264–268. [Google Scholar] [CrossRef]
  137. Zadorozhnyy, M.Y.; Chukov, D.I.; Churyukanova, M.N.; Gorshenkov, M.V.; Zadorozhnyy, V.Y.; Stepashkin, A.A.; Tsarkov, A.A.; Louzguine-Luzgin, D.V.; Kaloshkin, S.D. Investigation of contact surfaces between polymer matrix and metallic glasses in composite materials based on high-density polyethylene. Mater. Des. 2016, 92, 306–312. [Google Scholar] [CrossRef]
  138. Sharma, A.; Zadorozhnyy, M.; Stepashkin, A.; Kvaratskheliya, A.; Korol, A.; Moskovskikh, D.; Kaloskin, S.; Zadorozhnyy, V. Investigation of Thermophysical Properties of Zr-Based Metallic Glass-Polymer Composite. Metals 2021, 11, 1412. [Google Scholar] [CrossRef]
  139. Nowosielski, R.; Gramatyka, P.; Sakiewicz, P.; Babilas, R. Ferromagnetic composites with polymer matrix consisted of nanocrystalline Fe-based filler. J. Magn. Magn. Mater. 2015, 387, 179–185. [Google Scholar] [CrossRef]
  140. Gizaw, E.T.; Yeh, H.-H.; Chu, J.P.; Hu, C.-C. Fabrication and characterization of nitrogen selective thin-film metallic glass/polyacrylonitrile composite membrane for gas separation. Sep. Purif. Technol. 2020, 237, 116340. [Google Scholar] [CrossRef]
  141. Kim, J.-Y.; Gu, X.; Wraith, M.; Uhl, J.T.; Dahmen, K.A.; Greer, J.R. Suppression of Catastrophic Failure in Metallic Glass-Polyisoprene Nanolaminate Containing Nanopillars. Adv. Funct. Mater. 2012, 22, 1972–1980. [Google Scholar] [CrossRef]
  142. Lasheras, A.; Gutiérrez, J.; Reis, S.; Sousa, D.; Silva, M.; Martins, P.; Lanceros-Mendez, S.; Barandiarán, J.M.; Shishkin, D.A.; Potapov, A.P. Energy harvesting device based on a metallic glass/PVDF magnetoelectric laminated composite. Smart Mater. Struct. 2015, 24. [Google Scholar] [CrossRef]
  143. Eckert, J.; Das, J.; Pauly, S.; Duhamel, C. Mechanical properties of bulk metallic glasses and composites. J. Mater. Res. 2007, 22, 285–301. [Google Scholar] [CrossRef]
  144. Lin, S.; Ge, S.; Li, W.; Li, H.; Fu, H.; Wang, A.; Zhuang, Y.; Zhang, H.; Zhu, Z. Work-hardenable TiZr-based multilayered bulk metallic glass composites through the solid solution strengthening in ex-situ Ti layers. J. Non-Cryst. Solids 2021, 553, 120508. [Google Scholar] [CrossRef]
  145. Lin, S.; Zhu, Z.; Ge, S.; Zhang, L.; Liu, D.; Zhuang, Y.; Fu, H.; Li, H.; Wang, A.; Zhang, H. Designing new work-hardenable ductile Ti-based multilayered bulk metallic glass composites with ex-situ and in-situ hybrid strategy. J. Mater. Sci. Technol. 2020, 50, 128–138. [Google Scholar] [CrossRef]
  146. Wang, X.; Zhao, L.; Hu, X.; Cheng, Y.; Liu, S.; Chen, P.; Cui, C. Fabrication and Mechanical Behavior of Ex Situ Mg-Based Bulk Metallic Glass Matrix Composite Reinforced with Electroless Cu-Coated SiC Particles. Materials 2017, 10, 1371. [Google Scholar] [CrossRef] [Green Version]
  147. Qin, Y.; Han, X.; Song, K.; Wang, L.; Cheng, Y.; Zhang, Z.; Xue, Q.; Sun, N.; Wang, J.; Sun, B.; et al. Stability of shear banding process in bulk metallic glasses and composites. J. Mater. Res. 2017, 32, 2560–2569. [Google Scholar] [CrossRef]
  148. Guo, W.; Wada, T.; Kato, H. Work-hardenable Mg-based bulk metallic glass matrix composites reinforced by ex-situ porous shape-memory-alloy particles. Mater. Lett. 2016, 183, 454–458. [Google Scholar] [CrossRef]
  149. Li, T.H.; Tsai, P.H.; Hsu, K.T.; Liu, Y.C.; Jang, J.S.C.; Huang, J.C. Significantly enhanced drilling ability of the orthopedic drill made of Zr-based bulk metallic glass composite. Intermetallics 2016, 78, 17–20. [Google Scholar] [CrossRef]
  150. Wang, H.-S.; Wu, J.-Y.; Liu, Y.-T. Effect of the volume fraction of the ex-situ reinforced Ta additions on the microstructure and properties of laser-welded Zr-based bulk metallic glass composites. Intermetallics 2016, 68, 87–94. [Google Scholar] [CrossRef]
  151. Wang, H.-S.; Chen, H.-G.; Jang, J.S.-C.; Lin, D.-Y.; Gu, J.-W. Interfacial analysis of the ex-situ reinforced phase of a laser spot welded Zr-based bulk metallic glass composite. Mater. Charact. 2013, 86, 242–249. [Google Scholar] [CrossRef]
  152. Madge, S.V.; Louzguine-Luzgin, D.V.; Inoue, A.; Greer, A.L. Large Compressive Plasticity in a La-Based Glass-Crystal Composite. Metals 2013, 3, 41–48. [Google Scholar] [CrossRef]
  153. Jang, J.S.C.; Li, J.B.; Lee, S.L.; Chang, Y.S.; Jian, S.R.; Huang, J.C.; Nieh, T.G. Prominent plasticity of Mg-based bulk metallic glass composites by ex-situ spherical Ti particles. Intermetallics 2012, 30, 25–29. [Google Scholar] [CrossRef]
  154. Li, J.B.; Jang, J.S.C.; Li, C.; Jian, S.R.; Tsai, P.H.; Hwang, J.D.; Huang, J.C.; Nieh, T.G. Significant plasticity enhancement of ZrCu-based bulk metallic glass composite dispersed by in situ and ex situ Ta particles. Mater. Sci. Eng. A 2012, 551, 249–254. [Google Scholar] [CrossRef]
  155. Li, J.B.; Jang, J.S.C.; Jian, S.R.; Chen, K.W.; Lin, J.F.; Huang, J.C. Plasticity improvement of ZrCu-based bulk metallic glass by ex situ dispersed Ta particles. Mater. Sci. Eng. A 2011, 528, 8244–8248. [Google Scholar] [CrossRef]
  156. Qiao, J.; Jia, H.; Liaw, P.K. Metallic glass matrix composites. Mater. Sci. Eng. R: Rep. 2016, 100, 1–69. [Google Scholar] [CrossRef] [Green Version]
  157. Lin, S.; Ge, S.; Zhu, Z.; Li, W.; Li, Z.; Li, H.; Fu, H.; Wang, A.; Zhuang, Y.; Zhang, H. Double toughening Ti-based bulk metallic glass composite with high toughness, strength and tensile ductility via phase engineering. Appl. Mater. Today 2021, 22, 100944. [Google Scholar] [CrossRef]
  158. Ren, Z.Q.; Churakova, A.A.; Wang, X.; Goel, S.; Liu, S.N.; You, Z.S.; Liu, Y.; Lan, S.; Gunderov, D.V.; Wang, J.T.; et al. Enhanced tensile strength and ductility of bulk metallic glasses Zr52.5Cu17.9Al10Ni14.6Ti5 via high-pressure torsion. Mater. Sci. Eng. A 2020, 803, 140485. [Google Scholar] [CrossRef]
  159. Jin, Y.; Xi, G.; Li, R.; Li, Z.-A.; Chen, X.-B.; Zhang, T. Nanoporous metallic-glass electrocatalysts for highly efficient oxygen evolution reaction. J. Alloys Compd. 2021, 852, 156876. [Google Scholar] [CrossRef]
  160. Zhang, J.; Liu, Y.; Yang, H.; Ren, Y.; Cui, L.; Jiang, D.; Wu, Z.; Ma, Z.; Guo, F.; Bakhtiari, S.; et al. Achieving 5.9% elastic strain in kilograms of metallic glasses: Nanoscopic strain engineering goes macro. Mater. Today 2020, 37, 18–26. [Google Scholar] [CrossRef]
  161. Ibrahim, M.Z.; Sarhan, A.A.D.; Kuo, T.Y.; Yusof, F.; Hamdi, M.; Lee, T.M. Developing a new laser cladded FeCrMoCB metallic glass layer on nickel-free stainless-steel as a potential superior wear-resistant coating for joint replacement implants. Surf. Coat. Technol. 2020, 392, 125755. [Google Scholar] [CrossRef]
  162. Tkaczyk, M.; Hasiak, M.; Kaleta, J.; Dragnevski, K.I. Microstructural investigation and mechanical properties of rapidly solidified bulk nanocrystalline Fe-based alloys. Mater. Today: Proc. 2020, 33, 1775–1780. [Google Scholar] [CrossRef]
  163. Huang, L.J.; Wang, H.; Liu, J.W.; Zhang, C.; Ouyang, L.Z.; Zhu, M. Low temperature de/hydrogenation in the partially crystallized Mg60Ce10Ni20Cu10 metallic glasses induced by milling with process control agents. J. Alloys Compd. 2019, 792, 835–843. [Google Scholar] [CrossRef]
  164. Hasani, S.; Ansariniya, M.; Seifoddini, A. Enhancement of mechanical properties of a soft magnetic Fe-based metallic glass. Mater. Sci. Technol. 2019, 35, 865–871. [Google Scholar] [CrossRef]
  165. Zhou, Q.; Ren, Y.; Du, Y.; Han, W.; Hua, D.; Zhai, H.; Huang, P.; Wang, F.; Wang, H. Identifying the significance of Sn addition on the tribological performance of Ti-based bulk metallic glass composites. J. Alloys Compd. 2019, 780, 671–679. [Google Scholar] [CrossRef]
  166. Yao, A.; Yang, H.; Wang, J.-Q.; Xu, W.; Huo, J.; Li, R.-W.; Qiu, H.; Chen, M. Flexible supercapacitor electrodes fabricated by dealloying nanocrystallized Al-Ni-Co-Y-Cu metallic glasses. J. Alloys Compd. 2019, 772, 164–172. [Google Scholar] [CrossRef]
  167. Monson, T.C.; Zheng, B.; Delany, R.E.; Pearce, C.J.; Langlois, E.D.; Lepkowski, S.M.; Stevens, T.E.; Zhou, Y.; Atcitty, S.; Lavernia, E.J. Soft Magnetic Multilayered FeSiCrB-FexN Metallic Glass Composites Fabricated by Spark Plasma Sintering. IEEE Magn. Lett. 2019, 10, 1–5. [Google Scholar] [CrossRef]
  168. Codrean, C.; Cosma, C.; Toșa, D.; Buzdugan, D.; Dume, A.I. Composite Materials Fabricated of Amorphous and Nanocrystalline Metallic Powders. Mater. Plast. 2019, 56, 744–749. [Google Scholar] [CrossRef]
  169. Ji, X.; Wu, J.; Pi, J.; Cheng, J.; Shan, Y.; Zhang, Y. Slurry erosion induced surface nanocrystallization of bulk metallic glass. Appl. Surf. Sci. 2018, 440, 1204–1210. [Google Scholar] [CrossRef]
  170. Hoppe, M.; Ababii, N.; Postica, V.; Lupan, O.; Polonskyi, O.; Schütt, F.; Kaps, S.; Sukhodub, L.F.; Sontea, V.; Strunskus, T.; et al. (CuO-Cu2O)/ZnO:Al heterojunctions for volatile organic compound detection. Sens. Actuators B: Chem. 2018, 255, 1362–1375. [Google Scholar] [CrossRef]
  171. Jian, W.-R.; Wang, L.; Yao, X.H.; Luo, S.-N. Balancing strength, hardness and ductility of Cu64Zr36 nanoglasses via embedded nanocrystals. Nanotechnology 2018, 29, 025701. [Google Scholar] [CrossRef] [PubMed]
  172. Tlili, A.; Pailhès, S.; Debord, R.; Ruta, B.; Gravier, S.; Blandin, J.-J.; Blanchard, N.; Gomès, S.; Assy, A.; Tanguy, A.; et al. Thermal transport properties in amorphous/nanocrystalline metallic composites: A microscopic insight. Acta Mater. 2017, 136, 425–435. [Google Scholar] [CrossRef]
  173. Kim, Y.S.; Hong, S.H.; Park, H.J.; Kim, J.T.; Jeong, H.J.; Na, Y.S.; Lim, K.R.; Park, J.M.; Kim, K.B. Crystallization and phase transformation behavior of TiCu-based bulk metallic glass composite with B2 particles. J. Alloys Compd. 2017, 707, 87–91. [Google Scholar] [CrossRef]
  174. Dusoe, K.J.; Vijayan, S.; Bissell, T.R.; Chen, J.; Morley, J.E.; Valencia, L.; Dongare, A.M.; Aindow, M.; Lee, S.-W. Strong, ductile, and thermally stable Cu-based metal-intermetallic nanostructured composites. Sci. Rep. 2017, 7, 40409. [Google Scholar] [CrossRef]
  175. Lan, X.; Wu, H.; Liu, Y.; Zhang, W.; Li, R.; Chen, S.; Zai, X.; Hu, T. Microstructures and tribological properties of laser cladded Ti-based metallic glass composite coatings. Mater. Charact. 2016, 120, 82–89. [Google Scholar] [CrossRef]
  176. Lee, J.-K.; Kim, S.-Y.; Ott, R.T.; Kim, J.-Y.; Eckert, J.; Lee, M.-H. Effect of reinforcement phase on the mechanical property of tungsten nanocomposite synthesized by spark plasma sintering. Int. J. Refract. Met. Hard Mater. 2016, 54, 14–18. [Google Scholar] [CrossRef] [Green Version]
  177. Kim, J.; Kyeong, J.S.; Ham, M.-H.; Minor, A.M.; Kim, D.H.; Park, E.S. Development of Mo-Ni-Si-B metallic glass with high thermal stability and H versus E ratios. Mater. Des. 2016, 98, 31–40. [Google Scholar] [CrossRef] [Green Version]
  178. Drescher, P.; Witte, K.; Yang, B.; Steuer, R.; Kessler, O.; Burkel, E.; Schick, C.; Seitz, H. Composites of amorphous and nanocrystalline Zr–Cu–Al–Nb bulk materials synthesized by spark plasma sintering. J. Alloys Compd. 2016, 667, 109–114. [Google Scholar] [CrossRef]
  179. Veligatla, M.; Das, S.; Lee, W.K.; Hwang, J.; Thumthan, O.; Hao, Y.; Mukherjee, S. Tuning the Magnetic Properties of Cobalt-Based Metallic Glass Nanocomposites. JOM 2016, 68, 336–340. [Google Scholar] [CrossRef]
  180. Xing, Q.; Zhang, K.; Wang, Y.; Leng, J.; Jia, H.; Liaw, P.K.; Wang, Y. Effects of pre-compression on the microstructure, mechanical properties and corrosion resistance of RE65Co25Al10 (RE = Ce, La, Pr, Sm and Gd) bulk metallic glasses. Intermetallics 2015, 67, 94–101. [Google Scholar] [CrossRef]
  181. Bai, J.; Wang, J.; Li, L.; Kou, H.; Li, J. Thermal Stability and the Matrix Induced Brittleness in a Ti-based Bulk Metallic Glass Composite. Mater. Res. 2015, 18, 83–88. [Google Scholar] [CrossRef] [Green Version]
  182. Babilas, R.; Nowosielski, R.; Pawlyta, M.; Fitch, A.; Burian, A. Microstructural characterization of Mg-based bulk metallic glass and nanocomposite. Mater. Charact. 2015, 102, 156–164. [Google Scholar] [CrossRef]
  183. Kuan, S.Y.; Huang, J.C. Improving the ductility of thin film metallic glasses via nano-twinning. Thin Solid Film. 2014, 561, 43–47. [Google Scholar] [CrossRef]
  184. Abdi, S.; Khoshkhoo, M.S.; Shuleshova, O.; Bönisch, M.; Calin, M.; Schultz, L.; Eckert, J.; Baró, M.D.; Sort, J.; Gebert, A. Effect of Nb addition on microstructure evolution and nanomechanical properties of a glass-forming Ti-Zr-Si alloy. Intermetallics 2014, 46, 156–163. [Google Scholar] [CrossRef]
  185. Zhu, S.; Xie, G.; Wang, H.; Yang, X.; Cui, Z.; Inoue, A. Zr-based bulk metallic glass composite with in situ precipitated nanocrystals. J. Alloys Compd. 2014, 586, 155–158. [Google Scholar] [CrossRef]
  186. Zou, L.M.; Li, Y.H.; Yang, C.; Qu, S.G.; Li, Y.Y. Effect of Fe content on glass-forming ability and crystallization behavior of a (Ti69.7Nb23.7Zr4.9Ta1.7)100−xFex alloy synthesized by mechanical alloying. J. Alloys Compd. 2013, 553, 40–47. [Google Scholar] [CrossRef]
  187. Wang, D.J.; Huang, Y.J.; Wu, L.Z.; Shen, J. Mechanical, acoustic and electrical properties of porous Ti-based metallic glassy/nanocrystalline composites. Mater. Des. 2013, 44, 69–73. [Google Scholar] [CrossRef]
  188. Zadorozhnyy, V.Y.; Inoue, A.; Louzguine-Luzgin, D.V. Formation and investigation of the structure and mechanical properties of bulk metallic glassy composite (Ti-Zr)-(Cu-Ni-Co) alloys. Intermetallics 2012, 31, 173–176. [Google Scholar] [CrossRef]
  189. Zadorozhnyy, V.Y.; Inoue, A.; Louzguine-Luzgin, D.V. Formation and investigation of the structure and mechanical properties of bulk metallic glassy composite (Ti-Zr)-(Cu-Ni-Co) alloys with the addition of Boron. Mater. Sci. Eng. A 2012, 558, 472–477. [Google Scholar] [CrossRef]
  190. Churyumov, A.Y.; Bazlov, A.; Solonin, A.N.; Zadorozhnyi, V.Y.; Xie, G.Q.; Li, S.; Louzguine-Luzgin, D.V. Structure and mechanical properties of Ni-Cu-Ti-Zr composite materials with amorphous phase. Phys. Met. Metallogr. 2013, 114, 773–778. [Google Scholar] [CrossRef]
  191. Zadorozhnyy, V.Y.; Inoue, A.; Louzguine-Luzgin, D.V. Investigation of the structure and mechanical properties of as-cast Ti-Cu-based alloys. Mater. Sci. Eng. A 2013, 573, 175–182. [Google Scholar] [CrossRef]
  192. Golovin, I.S.; Zadorozhnyy, V.Y.; Churyumov, A.Y.; Louzguine-Luzgin, D.; Louzguine-Luzgin, D. Internal friction in a Ni–Ti-based glassy-crystal alloy. J. Alloys Compd. 2013, 579, 633–637. [Google Scholar] [CrossRef]
  193. Tsarkov, A.A.; Churyumov, A.Y.; Zadorozhnyy, V.Y.; Louzguine-Luzgin, D.V. High-strength and ductile (Ti-Ni)-(Cu-Zr) crystalline/amorphous composite materials with superelasticity and TRIP effect. J. Alloys Compd. 2016, 658, 402–407. [Google Scholar] [CrossRef]
  194. Sergiienko, R.A.; Shcheretskyi, O.A.; Zadorozhnyy, V.Y.; Verkhovliuk, A.M.; Louzguine-Luzgin, D.V. Investigation of Zr55Cu30Al10Ni5 bulk amorphous alloy crystallization. J. Alloys Compd. 2019, 791, 477–482. [Google Scholar] [CrossRef]
  195. Tsarkov, A.A.; Zadorozhnyy, V.Y.; Solonin, A.N.; Louzguine-Luzgin, D.V. Effect of Aluminum, Iron and Chromium Alloying on the Structure and Mechanical Properties of (Ti-Ni)-(Cu-Zr) Crystalline/Amorphous Composite Materials. Metals 2020, 10, 874. [Google Scholar] [CrossRef]
  196. Duan, G.H.; Jiang, M.Q.; Liu, X.F.; Dai, L.H.; Li, J.X. In situ observations on shear-banding process during tension of a Zr-based bulk metallic glass composite with dendrites. J. Non-Cryst. Solids 2021, 565, 120841. [Google Scholar] [CrossRef]
  197. Bordeenithikasem, P.; Hofmann, D.C.; Firdosy, S.; Ury, N.; Vogli, E.; East, D.R. Controlling microstructure of FeCrMoBC amorphous metal matrix composites via laser directed energy deposition. J. Alloys Compd. 2021, 857, 157537. [Google Scholar] [CrossRef]
  198. Lee, J.I.; Wat, A.; Kim, J.; Ryu, C.W.; Chang, H.J.; Park, E.S.; Ritchie, R.O. Synthesis of bioinspired ice-templated bulk metallic glass-alumina composites with intertwined dendritic structure. Scr. Mater. 2019, 172, 159–164. [Google Scholar] [CrossRef] [Green Version]
  199. Li, L.; Li, J.; He, Y.; Lai, M.; Zhu, M.; Wang, W.Y.; Kou, H.; Wang, J. A new microscopic, coordinated deformation model of Ti-based bulk metallic composites during tensile deformation. Scr. Mater. 2019, 172, 23–27. [Google Scholar] [CrossRef]
  200. Zhang, L.; Narayan, R.L.; Fu, H.M.; Ramamurty, U.; Li, W.R.; Li, Y.D.; Zhang, H.F. Tuning the microstructure and metastability of β-Ti for simultaneous enhancement of strength and ductility of Ti-based bulk metallic glass composites. Acta Mater. 2019, 168, 24–36. [Google Scholar] [CrossRef]
  201. Wang, Y.S.; Tian, F.; Qiao, B.B.; Ma, S.M.; Wang, X.T.; Wang, X.M.; Lan, A.D. Mechanical property and serration behavior of Ti-based metallic glassy composites reinforced by an in situ dendritic phase. Mater. Sci. Eng. A 2019, 743, 301–308. [Google Scholar] [CrossRef]
  202. Yang, S.; Li, X.C.; Wei, R.; Meng, M.; He, L.; Zhang, S.F. Deformation characteristic of a Ti-based bulk metallic glass composite with fine microstructure. Mater. Sci. Eng. A 2018, 733, 224–231. [Google Scholar] [CrossRef]
  203. Kim, W.C.; Kim, K.C.; Na, M.Y.; Jeong, S.H.; Kim, W.T.; Kim, D.H. Formation of crystalline phase in the glass matrix of Zr-Co-Al glass-matrix composites and its effect on their mechanical properties. Met. Mater. Int. 2017, 23, 1216–1222. [Google Scholar] [CrossRef]
  204. Hua, N.; Yue, F. Ultrafine-structured Ni-based bulk alloys with high strength and enhanced ductility. Mater. Sci. Eng. A 2017, 692, 17–23. [Google Scholar] [CrossRef]
  205. Okulov, I.V.; Bönisch, M.; Volegov, A.S.; Shahabi, H.S.; Wendrock, H.; Gemming, T.; Eckert, J. Micro-to-nano-scale deformation mechanism of a Ti-based dendritic-ultrafine eutectic alloy exhibiting large tensile ductility. Mater. Sci. Eng. A 2017, 682, 673–678. [Google Scholar] [CrossRef]
  206. Bu, F.; Wang, J.; Li, L.; Kou, H.; Xue, X.; Li, J. The Effect of Thermal Cycling Treatments on the Thermal Stability and Mechanical Properties of a Ti-Based Bulk Metallic Glass Composite. Metals 2016, 6, 274. [Google Scholar] [CrossRef] [Green Version]
  207. Yang, S.; Li, D.; Wang, X.S.; Guo, J.W.; Zhang, S.F.; He, L. Mechanical behavior and wear performance of a Ti-based bulk metallic glass composite containing dendritic and intermetallic phases. Mater. Sci. Eng. A 2016, 672, 135–142. [Google Scholar] [CrossRef]
  208. Gargarella, P.; Pauly, S.; Khoshkhoo, M.S.; Kiminami, C.S.; Kühn, U.; Eckert, J. Improving the glass-forming ability and plasticity of a TiCu-based bulk metallic glass composite by minor additions of Si. J. Alloys Compd. 2016, 663, 531–539. [Google Scholar] [CrossRef]
  209. Cui, J.; Li, J.S.; Wang, J.; Kou, H.C. Microstructure Evolution and Mechanical Properties of a Ti-Based Bulk Metallic Glass Composite. J. Mater. Eng. Perform. 2015, 24, 2354–2358. [Google Scholar] [CrossRef]
  210. Li, Y.L.; Kou, H.; Wang, J.; Li, J.; Hu, R. Enhanced mechanical properties of Ti-based metallic glass composites prepared under medium vacuum system. J. Non-Cryst. Solids 2015, 413, 15–19. [Google Scholar] [CrossRef]
  211. Ma, D.Q.; Jiao, W.T.; Zhang, Y.F.; Wang, B.A.; Li, J.; Zhang, X.Y.; Ma, M.Z.; Liu, R.P. Strong work-hardening behavior induced by the solid solution strengthening of dendrites in TiZr-based bulk metallic glass matrix composites. J. Alloys Compd. 2015, 624, 9–16. [Google Scholar] [CrossRef]
  212. Okulov, I.; Wendrock, H.; Volegov, A.; Attar, H.; Kühn, U.; Skrotzki, W.; Eckert, J. High strength beta titanium alloys: New design approach. Mater. Sci. Eng. A 2015, 628, 297–302. [Google Scholar] [CrossRef]
  213. Arora, H.S.; Mridha, S.; Grewal, H.S.; Singh, H.; Hofmann, D.C.; Mukherjee, S. Controlling the length scale and distribution of the ductile phase in metallic glass composites through friction stir processing. Sci. Technol. Adv. Mater. 2014, 15, 035011. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Chen, J.H.; Jiang, M.Q.; Chen, Y.; Dai, L.H. Strain rate dependent shear banding behavior of a Zr-based bulk metallic glass composite. Mater. Sci. Eng. A 2013, 576, 134–139. [Google Scholar] [CrossRef] [Green Version]
  215. Qiao, J.-W.; Zhang, Y.; Yang, H.-J.; Sang, S.-B. Dendritic and spherical crystal reinforced metallic glass matrix composites. Int. J. Miner. Metall. Mater. 2013, 20, 386–392. [Google Scholar] [CrossRef]
  216. Antonysamy, A.A.; Pauly, S.; Dhindaw, B.K.; Eckert, J. Influence of Superheat on Microstructure and Mechanical Properties of Ductile Cu47.5Zr47.5Al5 Bulk Metallic Glass-Matrix Composite. J. Mater. Eng. Perform. 2011, 20, 1196–1205. [Google Scholar] [CrossRef]
  217. Liu, B.-C.; Zhang, Q.-D.; Wang, H.-J.; Li, X.-Y.; Zu, F.-Q. Significantly improved plasticity of bulk metallic glasses by introducing quasicrystal within high energy glass matrix. Intermetallics 2019, 111, 106504. [Google Scholar] [CrossRef]
  218. Zhao, Y.-Y.; Men, H.; Estevez, D.; Liu, Y.; Wang, X.; Li, R.-W.; Chang, C. Mg-based bulk metallic glass composite containing in situ microsized quasicrystalline particles. Scr. Mater. 2014, 78-79, 21–24. [Google Scholar] [CrossRef]
  219. Zhang, J.L.; Wang, Y.M.; Shek, C.H. Correlation between structures and properties in (Zr65Cu15Ni10Al10)90Nb10 alloys. J. Mater. Res. 2013, 28, 1218–1223. [Google Scholar] [CrossRef]
Figure 1. The elastic limit (MPa) plotted against Young’s modulus (GPa) for different materials. In total, 1507 metals, metal–matrix composites, alloys, and metallic glasses (at%) are compared. (σy/E) contour represents elastic strain limit and (σy2/E) as resilience. Reproduced from the permission of ref. [53]. Copyright 2009 Elsevier Ltd.
Figure 1. The elastic limit (MPa) plotted against Young’s modulus (GPa) for different materials. In total, 1507 metals, metal–matrix composites, alloys, and metallic glasses (at%) are compared. (σy/E) contour represents elastic strain limit and (σy2/E) as resilience. Reproduced from the permission of ref. [53]. Copyright 2009 Elsevier Ltd.
Metals 11 01933 g001
Figure 2. The bio-metallic glass is formed by bioglass and crystalline bio-metallic alloys. This also shows the relationship between bioglass, bio-metallic alloys, and bio-metallic glass. Reproduced with permission from ref. [71]. Copyright 2016 Elsevier Ltd.
Figure 2. The bio-metallic glass is formed by bioglass and crystalline bio-metallic alloys. This also shows the relationship between bioglass, bio-metallic alloys, and bio-metallic glass. Reproduced with permission from ref. [71]. Copyright 2016 Elsevier Ltd.
Metals 11 01933 g002
Figure 3. Hemocompatibility tests for different TFMGs (Ti-based thin film metallic glasses). (a) Control uncoated TFMGs, (b) TiB-0 (without Boron), (c) TiB-1 (4% B), (d) TiB-2 (8% B) and (e) TiB-3(14% B). Reproduced with permission from ref. [78]. Copyright 2021 Elsevier B.V.
Figure 3. Hemocompatibility tests for different TFMGs (Ti-based thin film metallic glasses). (a) Control uncoated TFMGs, (b) TiB-0 (without Boron), (c) TiB-1 (4% B), (d) TiB-2 (8% B) and (e) TiB-3(14% B). Reproduced with permission from ref. [78]. Copyright 2021 Elsevier B.V.
Metals 11 01933 g003
Figure 4. SEM images of tribocorrosion wear of (a) c-BMG, (b) a-BMG, and (c) Zr and alumina counterparts, (d) c-BMG, (e) a-BMG, and (f) Zr. Reproduced with permission from ref. [94]. Copyright © 2021 Elsevier B.V.
Figure 4. SEM images of tribocorrosion wear of (a) c-BMG, (b) a-BMG, and (c) Zr and alumina counterparts, (d) c-BMG, (e) a-BMG, and (f) Zr. Reproduced with permission from ref. [94]. Copyright © 2021 Elsevier B.V.
Metals 11 01933 g004
Figure 5. A comparison of corrosion rates of the Mg-Zn-Ca-Sr (BMGs), biodegradable BMGs, and crystalline Mg alloys in physiological solution. Reproduced with permission from ref. [104]. Copyright 2014 Elsevier Ltd.
Figure 5. A comparison of corrosion rates of the Mg-Zn-Ca-Sr (BMGs), biodegradable BMGs, and crystalline Mg alloys in physiological solution. Reproduced with permission from ref. [104]. Copyright 2014 Elsevier Ltd.
Metals 11 01933 g005
Figure 6. Morphologies of NIH 3T3 cells (Swiss mouse embryo fibroblasts 3-day transfer, inoculum 3 × 105 cells) cultured on (a) 316L SS, (b) TC4, (c) Cr10Mo10, (d) Cr20Mo5 and (e) Cr20Mo5 for 2 days. On the right side are enlarged micrographs of the counterparts. Reproduced with permission from ref. [108]. Copyright 2015 Elsevier B.V.
Figure 6. Morphologies of NIH 3T3 cells (Swiss mouse embryo fibroblasts 3-day transfer, inoculum 3 × 105 cells) cultured on (a) 316L SS, (b) TC4, (c) Cr10Mo10, (d) Cr20Mo5 and (e) Cr20Mo5 for 2 days. On the right side are enlarged micrographs of the counterparts. Reproduced with permission from ref. [108]. Copyright 2015 Elsevier B.V.
Metals 11 01933 g006
Figure 7. A representation of the (a) new continuous infiltration process, (b) a bunch of BMG MC (bulk metallic glass matrix composites), (c,d) SEM images of the cross section of BMG MC bunch having 61.4 vol% tungsten wires and (e) TEM image of the interaction between the tungsten and the Zr41.25Ti13.75Cu12.5Ni10Be22.5 (Vit 1, at%) shown in the inset of (d). Reproduced with permission from ref. [129]. Copyright 2012 Elsevier B.V.
Figure 7. A representation of the (a) new continuous infiltration process, (b) a bunch of BMG MC (bulk metallic glass matrix composites), (c,d) SEM images of the cross section of BMG MC bunch having 61.4 vol% tungsten wires and (e) TEM image of the interaction between the tungsten and the Zr41.25Ti13.75Cu12.5Ni10Be22.5 (Vit 1, at%) shown in the inset of (d). Reproduced with permission from ref. [129]. Copyright 2012 Elsevier B.V.
Metals 11 01933 g007
Figure 8. SEM images of fracture area of (a) monolithic Zr47.3Cu36Ag8Al8Si0.7 BMG, (b) Zr47.3Cu32Ag8Al8Ta4Si0.7 BMGC, (c) Zr47.3Cu32Ag8Al8Ta4Si0.7BMGC with ex situ added, 3 vol.% Ta, and (d) Zr47.3Cu32Ag8Al8Ta4Si0.7 BMGC with ex situ added, 9 vol.% Ta. Reproduced with permission from ref. [154]. Copyright 2012 Elsevier B.V.
Figure 8. SEM images of fracture area of (a) monolithic Zr47.3Cu36Ag8Al8Si0.7 BMG, (b) Zr47.3Cu32Ag8Al8Ta4Si0.7 BMGC, (c) Zr47.3Cu32Ag8Al8Ta4Si0.7BMGC with ex situ added, 3 vol.% Ta, and (d) Zr47.3Cu32Ag8Al8Ta4Si0.7 BMGC with ex situ added, 9 vol.% Ta. Reproduced with permission from ref. [154]. Copyright 2012 Elsevier B.V.
Metals 11 01933 g008
Figure 9. XRD pattern of (a) DSC curve, (b) SEM-BSE (backscattered electron) image, (c) and TEM BF image, (d) of as-cast alloy, together with SAED patterns obtained from an amorphous matrix, (e,f) B2 particle. Reproduced with permission from ref. [173]. Copyright 2016 Elsevier B.V.
Figure 9. XRD pattern of (a) DSC curve, (b) SEM-BSE (backscattered electron) image, (c) and TEM BF image, (d) of as-cast alloy, together with SAED patterns obtained from an amorphous matrix, (e,f) B2 particle. Reproduced with permission from ref. [173]. Copyright 2016 Elsevier B.V.
Metals 11 01933 g009
Figure 10. Compressive engineering stress–strain (σε) curves of the base and Nb-added alloys, Reproduced with permission from ref. [207]. Copyright 2016 Elsevier B.V.
Figure 10. Compressive engineering stress–strain (σε) curves of the base and Nb-added alloys, Reproduced with permission from ref. [207]. Copyright 2016 Elsevier B.V.
Metals 11 01933 g010
Table 1. Different critical cooling rates (Rc) of metallic glass.
Table 1. Different critical cooling rates (Rc) of metallic glass.
Metallic GlassRc (K s−1)References
Au55Cu25Si203.4 × 104[41]
Fe48Cr15Mo14Y2C15B6140–190[42]
Co75Si15B103.8 × 108[43]
Fe79Si10B113.7 × 108[43]
Ni75Si8B172.4 × 108[43]
Pd77.5Cu6Si16.51.5 × 108[43]
Pd40Ni40P201.4 × 107[43]
Zr41.2Ti13.8Cu12.5Ni10.0Be22.5≤10[44]
Al51Ge35Ni14104[45]
Fe76Si9B10P5≤550[46]
Pd40Cu30Ni10P202.08 × 10−5[47]
Au49Ag5.5Pd2.3Cu26.9Si16.3600[48]
Zr52.5Ti5Al10Ni14.6Cu17.9103–105[49]
Fe67Mo4.5Cr2.3Al2Si3C7P8.7B5.59.3 × 104[50]
Al86Ni6Y4.5Co2La1.53.01 × 103[51]
Table 2. Mechanical and corrosion properties of the Ti-based metallic glass and composite.
Table 2. Mechanical and corrosion properties of the Ti-based metallic glass and composite.
Chemical CompositionPreparation MethodYoung’s Modulus (GPa)Compressive strength (MPa)Corrosion Values Ecorr(V)/Icorr (A·m−2)Reference
Ti45Zr10Cu31Pd10Sn4Gas atomization100 ± 3.22060 ± 85Ecorr = −0.206[75]
Ti45Zr10Cu31Pd10Sn4Spark Plasma sintering10–302060 [76]
Ti51Zr5Cu41Sn3Laser Cladding Icorr = 10−5[77]
Ti45Zr5Cu41Ni6Sn3Pulsed laser depositionIcorr = 1.33[78]
Ti45Zr20Cu35 (high-Cu)Ti45Zr20Cu35 (high-Cu)77Ecorr = –0.259 ± 0.041[79]
Ti47Cu38−xZr7.5Fe2.5Sn2Si1Ag2Tax (x = 1–4)Copper mold casting1002028 ± 32[80]
Ti47Cu38Zr7.5Fe2.5Sn2Si1Nb2Laser surface remelting153.6Icorr = 12 × 10−6[81]
Ti47Cu38−xZr7.5Fe2.5Sn2Si1Ag2Nbx (x = 0, 1,2; at%)Melt spinning Process 100.4 ± 0.12010 ± 66Icorr = 0.1 Am−2[82]
Ti40Zr10Cu36Pd14Tilt copper mold-casting961930Icorr = 6 × 10−9[83]
Ti60Zr10Nb15Si15Melt spinningEcorr = −0.195 ± 0.025[84]
Ti41.5Zr2.5Hf5Cu37.5Ni7.5Si1Sn5Induction heating150.4 ± 4.72000 ± 78Icorr = 10−5[85]
Table 3. Mechanical and corrosion properties of the Zr-based metallic glass and composite.
Table 3. Mechanical and corrosion properties of the Zr-based metallic glass and composite.
Chemical CompositionPreparation MethodYoung’s Modulus (GPa)Compressive Strength (MPa)Corrosion Values
Ecorr(V)/Icorr (A·m−2)
Reference
Zr44Ti11Cu10Ni11Be2Ceramic conversion treatments145.70 ± 3.13Ecorr = −0.02 V[88]
Zr48Cu36Al8Ag8Magnetron sputtering113.92Icorr = 7.24 × 10−6[91]
Zr60 + xTi2.5Al10Fe12.5− xCu10Ag5 (at.%, x = 0, 2.5, 5)Copper mold casting70–78240–255 ± 3Icorr = 0.86–1.23[93]
Table 4. Mechanical and corrosion properties of the Mg-based metallic glass and composite.
Table 4. Mechanical and corrosion properties of the Mg-based metallic glass and composite.
Chemical CompositionPreparation MethodCompressive Strength (MPa)Corrosion Rates Ecorr (V)/Rcorr(w/sqt)Reference
Relaxed Mg70Zn26Ca4 ribbon/Crystallized Mg70Zn26Ca4 ribbonMelt spinningEcorr = (−1.21 ± 0.065/−1.38 ± 0.087[99]
Mg65Zn30Ca5Melt pinning/ball-on-disk reciprocating sliding715–854Ecorr = −1.23[100]
Mg66Zn30Ca4xSrx(x = 0, 0.5, 1, and 1.5 at.%)Induction-melting787 ± 22–847 ± 24Rcorr = 8.76 × 104[104]
Mg67Zn27Ca4Induction-melting545–364−1.116[106]
Table 5. Mechanical properties of fiber-reinforced metallic glass composite.
Table 5. Mechanical properties of fiber-reinforced metallic glass composite.
Chemical CompositionPreparation MethodYield Strength (MPa)Fracture Strength (MPa)Plastic Strain (%)Reference
Carbon fiber/Ti37.3Zr22.7Be25.5Fe5.5Cu9Vacuum pressure infiltration processing1880 ± 804.8 ± 1.0%[113]
Tungsten fiber/Zr41.2Ti13.8Cu10Ni12.5Be22.5Forward melt infiltrating method1600 11.7%[114]
Tungsten fiber/(Zr40.08Ti13.30Cu11.84Ni10.07Be24.71)99Nb1Quasi-static compression2200–250026%[115]
TiNi fiber/Zr47Ti13Cu11Ni10-Be16Nb3Pressure infiltration casting.210014%[116]
Tungsten fiber/Zr41.2Ti13.8Ni10.0Cu12.5Be22.5Infiltration and rapid solidification37%[117]
Tungsten strings/(Zr40.08Ti13.30Ni11.84Cu10.07Be24.71)99NbInfiltration and rapid solidification method1680250527.1%[118]
Tungsten fiber/Zr41.2Ti13.8Cu12.5Ni10.0Be22.5Resistive furnace and quenching210020%[120]
Al matrix composites/Fe74Mo4P10C7.5B2.5Si2(at.%) metallic glass fibersBall milling/Hot pressing115 ± 8420040%[122]
Tungsten fiber/Zr41.2Ti13.8Cu12.5Ni10Be22.5Infiltration and rapid solidification23%[124]
Tungsten fiber/Zr41.2Ti13.8Ni12.5Cu10Be22.5Infiltration and rapid solidification method43312.6%[127]
Tungsten fiber/Zr41.25Ti13.75Cu12.5Ni10Be22.5Continuous infiltration process12300.75%[128]
Wf(60% vol)/Zr41.25Ti13.75Cu12.5Ni10Be22.5Melt infiltration casting230024%[130]
Wf/(Zr41.2Ti13.8Cu12.5Ni10Be22.5)100-xNbx composites with x = 1, 3, 5 and 7.Melt infiltration casting245020%[132]
Table 6. Mechanical properties of ex situ metallic glass composite.
Table 6. Mechanical properties of ex situ metallic glass composite.
Chemical CompositionPreparation MethodYield Strength (MPa)Fracture Strength (MPa)Plastic Strain (%)Reference
Ti63 − xZrxNi5.3Cu9 Be22.7 (at.%, x = 52)Melt spinning/Ex situ method1552 ± 201956 ± 208.4 ± 0.2%[144]
(Ti0.328Zr0.302Ni0.053Cu0.09Be0.227)100 − xNbxMelt spinning/Quenching1078 ± 101112 ± 55.10 ± 0[145]
20%TiNi/Mg59.5Cu22.9Gd11Ag6.6Ex situ117312.1%[148]
La55Al25Cu10Ni10/35 mesh Ta particlesMelt spinning/Injection cast melting72040%[152]
Mg58Cu28.5Gd11Ag2.5/Ti particlesInjection casting 80025%[153]
Zr47.3Cu32Ag8Al8Ta4Si0.7/ex situ Ta particlesTwo-step arc melting process and suction casting1770–1800180044%[154]
Zr48Cu36Al8Ag8)99.25Si0.75/Ta particlesSuction casting1800185022%[155]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sharma, A.; Zadorozhnyy, V. Review of the Recent Development in Metallic Glass and Its Composites. Metals 2021, 11, 1933. https://0-doi-org.brum.beds.ac.uk/10.3390/met11121933

AMA Style

Sharma A, Zadorozhnyy V. Review of the Recent Development in Metallic Glass and Its Composites. Metals. 2021; 11(12):1933. https://0-doi-org.brum.beds.ac.uk/10.3390/met11121933

Chicago/Turabian Style

Sharma, Adit, and Vladislav Zadorozhnyy. 2021. "Review of the Recent Development in Metallic Glass and Its Composites" Metals 11, no. 12: 1933. https://0-doi-org.brum.beds.ac.uk/10.3390/met11121933

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