Next Article in Journal
The Influence of Alumina Airborne-Particle Abrasion on the Properties of Zirconia-Based Dental Ceramics (3Y-TZP)
Next Article in Special Issue
Isothermal Oxidation and Thermal Shock Resistance of Thick and Porous LaMgAl11O19 Abradable Topcoat
Previous Article in Journal
Stress Corrosion Cracking of Copper–Nickel Alloys: A Review
Previous Article in Special Issue
Properties and Performance of TiAlSiN and AlCrN Monolayer and Multilayer Coatings for Turning Ti-6Al-4V
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Distinctive Features and Fabrication Routes of Metallic-Glass Systems Designed for Different Engineering Applications: A Review

1
Department of Mechanical Engineering, The University of Texas at Dallas, Richardson, TX 75080, USA
2
Department of Materials Science and Engineering, University of Connecticut, Storrs, CT 06269, USA
3
School of Engineering & Technology, Central Michigan University, Mt. Pleasant, MI 48859, USA
4
The Science of Advanced Materials, Central Michigan University, Mt. Pleasant, MI 48859, USA
*
Author to whom correspondence should be addressed.
Coatings 2023, 13(10), 1689; https://doi.org/10.3390/coatings13101689
Submission received: 30 July 2023 / Revised: 10 September 2023 / Accepted: 12 September 2023 / Published: 26 September 2023
(This article belongs to the Special Issue Ceramic Materials and Coatings: From Materials to Applications)

Abstract

:
Materials with a disordered atomic structure, often termed glassy materials, are the focus of extensive research due to the possibility of achieving remarkable mechanical, electrochemical, and magnetic properties compared to crystalline materials. The glassy materials are observed to have an improved elastic modulus combined with a higher strength and hardness. Moreover, better corrosion resistance in different mediums is also observed for glassy solids, which is difficult to attain using conventional crystalline materials. As a result, the potential applications of metallic-glass systems are continually increasing. Amorphous materials are usually divided into two categories based upon their size. Materials with a thickness and diameter larger than the millimeter (mm) scale are termed as bulk metallic glass (BMG). However, the brittle nature of the bulk-sized samples restricts the size of metallic-glass systems to the micron (µm) or nanometer (nm) range. Metallic glasses with a specimen size in the scale of either µm or nm are defined as thin-film metallic glass (TFMG). In this review, BMGs and TFMGs are termed as metallic glass or MG. A large number of multi-component MGs and their compositional libraries reported by different research groups are summarized in this review. The formation of a multicomponent metallic glass depends on the constituent elements and the fabrication methods. To date, different unique fabrication routes have been adopted to fabricate BMG and TFMGs systems. An overview of the formation principles and fabrication methods as well as advantages and limitations of conventional MG fabrication techniques is also presented. Furthermore, an in-depth analysis of MG inherent properties, such as glass forming ability, and structural, mechanical, thermal, magnetic, and electrochemical properties, and a survey of their potential applications are also described.

1. Introduction

Metallic glasses (MG) are disordered non-equilibrium solids that, unlike crystalline alloys, lack long-range order. Metallic-glass systems are multicomponent with substantial differences in atomic sizes selected to suppress the crystallization of the liquid melt; extremely high cooling rates are used for vitrification [1]. Although metallic glasses possess high strength and a disordered structure like conventional glasses, high glass-transition temperatures are required for MGs [1]. Early work on metallic-glass systems started in the 1960s and 1970s [2,3,4]. The first disordered MG was a gold–silicon system developed by Jun et al. [2] and was fabricated via the melt cooling technique. Later on, Pd-Si-based metallic-glass systems (binary, ternary, and quaternary) with thicknesses of more than 1 mm were produced via the splat quenching technique under different cooling rates at room temperature by Chen et al. [3]. In 2000, metallic–glass systems (e.g., Fe-, Co- and Ni-based) were produced with thicknesses in the micrometer scale using cooling rates of ∼105 K/s [5].
Early metallic-glass systems were usually fabricated via the rapid quenching method to vitrify the liquid melt, thus requiring an extreme rate of cooling and introducing a size limitation on the metallic-glass systems [6]. However, Pd40Ni40P20 [7,8] and Pd-Cu-Si [4] glass systems are reported to have been fabricated with thicknesses of 10 mm using low cooling rates (∼10 K/s). These systems were found to have better glass-forming ability, but the high cost of palladium (Pd) hindered the practical implementation. The Zr-based (Zr-Ni-Al) metallic-glass systems reported by Inoue et al. [9] are considered to be the “earliest significant advancement” in glassy solids. Zr-based systems are observed to display excellent mechanical strength with a wide supercooled region, which can be useful for micro- or nano-fabrication [9,10]. The potential of Zr-based systems was further explored by Peker et al. [11], who included beryllium (Be) in Zr-Cu-Ni-Al. It is noteworthy that Be is the smallest atom in this metallic-glass system and its inclusion stabilizes the liquid and glassy phase [11]. This Be-containing amorphous alloy is a good illustration of the requirement for multiple species to form stable metallic glasses, chosen with a significant atomic size ratio.
The formation of metallic glasses also depends on the fabrication route. Solidification of the liquid melt through quenching is found to be helpful to maintain the glassy state; however, the very high cooling rate suggests an alternative technique. Conventional copper mold casting is found to be useful in this regard [6,12]. Lower quench rates of the order of 0.067–0.133 K/s have been reported for the fabrication of several metallic-glass systems [13], including Zr55Cu30Ni5Al10 [14], Pd40Cu30Ni10P20 [15], Ni50Pd30P20 [16], Mg59.5Cu22.9Ag6.6Gd11 [17], Cu36Zr48Ag8Al8 [18], Fe48Cr15Mo14Er2C15B6 [19], and Ti40Zr10Cu32Pd14Sn4 [20]. These systems are reported to have larger diameters (∼1–2 cm), which are useful in different engineering applications [12]. Nevertheless, the quench rates for the production of the amorphous solids still require optimization, and vapor-quenching techniques are found useful in this regard. Thin metallic-glass systems are produced using vapor-quenching techniques as reported in ref. [21,22,23,24]. The metallic glasses formed through the vapor quenching method are less brittle. Furthermore, the composition of the multicomponent systems also plays a vital part in glass formation. Deep eutectic points are found to be useful for obtaining amorphous structures, and choosing a composition near the eutectic point facilitates the formation of a metallic glass [25].
Metallic-glass systems are found to exhibit outstanding mechanical [26,27,28], electrochemical [29,30], and magnetic properties [31,32] due to the absence of dislocations and grain boundaries. In addition, using the glass-transition temperature of the metallic glass systems, they can be formed into unique structures by simple molding or drawing. [33,34]. Such outstanding features of the metallic-glass systems enable them to be used in biomedical systems [35,36], micro-electro-mechanical systems (MEMS) [37,38], nano-electro-mechanical systems [39,40], catalysis [41,42,43], and structural [44,45] applications. In this review, different fabrication techniques of the metallic-glass systems are described in combination with their unique properties and potential applications.

2. Glass Forming Ability

Glass forming ability (GFA) and its relationship with the atomic structure is crucial to the properties of metallic glasses. While discussing GFA, the “Confusion Principle” proposed by Lindsay Greer should be mentioned [46]. According to Greer, glass formation is facilitated due to increasing “confusion” during crystallization. The principle is primarily focused on the crystal structure and the atomic size of the components in multispecies systems [46]. Thus, the use of an additional element in the composition, such as a small beryllium atom, can facilitate metallic-glass formation [46]. The term “confusion” is further defined by Perim et al. [47], who states that impeding the growth of critical-size nuclei during synthesis can lead to the formation of an amorphous-state structure. According to this theory [47], the presence of multiple phases with different structures manifesting similar energies should form a metallic glass.
The empirical rules of metallic glass formation are further explained by Inoue et al. [5], who postulates three determining parameters for the possible formation of metallic glass: (i) the presence of multiple species, (ii) atomic size mismatch (≥12%) among the species, and (iii) the negative heat of mixing. The extensive classification of metallic-glass systems based on these parameters has also been reported Takeuchi et al. [48]. Moreover, compositional analyses of metallic-glass systems reveals that an optimum composition is required to achieve the maximum supercooled region. On the other hand, only 32 elements have been used from the periodic table for the analysis of metallic glass formation and a significant number of these combinations are produced through trial and error by different research groups [49]. For example, a study carried out by Zeman et al. [50] assessed the optimum composition for Zr-Cu binary systems by varying the copper content from 18 to 88 at.%. They found that a Cu content within the range 30–65 at.% facilitates glass transition; at 55 at.% of Cu the maximum supercooled region is achieved.
Deviation from the glass-forming rules of Inoue is often observed as the method of fabrication also plays a vital role in the formation of metallic glass. The most commonly reported metallic-glass formation techniques are liquid-to-solid transition methods, which require a high range of critical cooling rates [51,52,53]. However, a high cooling rate hampers the production of large-size metallic-glass systems and may contribute to partial crystallinity in the material matrix [54]. For example, Calin et al. [54] fabricated Ti75Zr10Si15 and Ti60Nb15Zr10Si15 systems via arc melting. The first system possessed a positive heat of mixing and a Ti-Zr atomic mismatch below 12%, and their results revealed a partially crystalline metallic glass structure. For the quaternary system, the Ti-Nb bond exhibited a positive heat of mixing, and the atomic mismatch also did not fulfil Inoue’s rule; in that case as well, the synthesis led to the formation of a metallic glass with partial crystallinity. These issues can be solved with a higher rate of vapor quenching [13]. Metallic-glass systems produced via different techniques are listed in Table 1 and the systems are assessed with respect to Inoue’s empirical rule.

3. Fabrication Techniques

Different fabrication routes can be adopted to provide randomness in the atomic orientation of the multicomponent systems. Metallic glasses are formed primarily by conventional casting or thermal-quenching techniques. Some novel fabrication techniques with their distinctive features will also be discussed in the following sections.

3.1. Liquid-Quenching Method

A significant fraction of bulk metallic-glass systems are formed by the liquid-to-solid transition technique [70,71,72]. To form a multicomponent metallic glass, metallic species are converted to a liquid by heating above the melting/liquidus temperature and later solidified by fast cooling. Thus, the liquid-quenching step requires a wide range of cooling rates. Based on the required cooling rate, the solidification technique can be classified into several groups, which are listed in Table 2. The extreme cooling rate in quenching techniques is found to be reached by distributing a thin layer of liquid in contact with a thermally conductive substrate (metal or sapphire) [25].
The quenching rate is defined with reference to the heat-transfer capacity between the liquid and the substrate as well as on the thickness and thermal conductivity of the liquid layer [25]. For example, to quench the liquid in a conventional die-casting method requires a cooling rate within the range of 10 K/s to 103 K/s [53]. Two temperature transitions occur in the formation of glass during quenching. The first transition of temperature is the liquidous temperature and the other one is the glass-transition temperature. According to Greer [74], decreasing the difference between these two temperatures favors the formation of metallic glass. In addition, glass formation is more likely to occur with rapid cooling of the liquid [74]. To extensively study the effect of the cooling rate, Kuball et al. [86] investigated the crystallization behavior of the Al86Ni8Y6 system. As illustrated in Figure 1, Kuball et al. [86] concluded that apart from the material composition, considerably high cooling rates can limit the formation and growth of the intermetallic compound (Al23Ni6Y4) and solid solution phase (α-Al) and form a monolithic metallic glass [86].
Although metallic-glass systems produced through rapid quenching techniques are typically multicomponent, a monoatomic metallic-glass system can be fabricated using extremely high cooling rates (1014 K/s) [87]. Zhong et al. brought two Ta nano-tips into contact and melted them using short square-wave electric pulses (∼3.7 ns and 0.5–3 V) [87]. As shown in Figure 2, they showed that the liquid was vitrified with a very high cooling rate to form a Ta metallic glass with a length and thickness on the nanometer scale [87]. However, the requirement of enormously high cooling rates to circumvent crystallization sets a severe restriction on the dimensions and geometry of the fabricated samples reachable using this process.

3.2. Welding of Metallic Glasses

The commercial production of metallic-glass systems is hampered by the requirement of a high cooling rate to fabricate a large-sized sample via casting. To eliminate the size constraint and fabricate bulk amorphous systems, a welding approach has been explored to join pieces of metallic glasses without compromising their excellent physical properties. The fabrication of a metallic glass composite (as shown in Figure 3) via this technique extends the application range of metallic-glass systems [88].
Different liquid-state joining techniques, such as electron-beam welding [89], laser-beam [88,90] welding, the pulse-current method [91], and gas tungsten arc welding [92] have been studied for different multicomponent metallic glasses. Liquid-state joining processes exhibit high energy density and deep weld penetration, which are helpful compared to conventional welding techniques [93]. However, phase stabilization in the heat-affected and weld-fusion zones to form a completely disordered solid remains a research challenge [94]. Among the liquid-state welding techniques, electron- and laser-beam welding have been extensively investigated. For example, a study conducted by Wang et al. [94] on the formation of Ti-based (Ti40Zr25Ni3Cu12Be20) metallic glass using laser welding showed that the microstructure of both the weld fusion zone (WFZ) and heat-affected zone (HAZ) are affected by the welding speed. In the WFZ, a high-speed welding of 10 m/min provides a high cooling rate of 780 K/s, which results in an amorphous microstructure. In the WFZ, the lower welding speed would cause a high temperature exposure for a longer period of time, leading to crystallization in the microstructure. However, at higher welding speed, the HAZ would experience a fast temperature drop below the crystallization temperature, suppressing the crystallization [94]. The reason for this is attributed to the high heating rate; decreasing the welding speed tends to suppress crystallization [94].
Solid-state joining techniques, including friction [95], explosion [96], thermoplastic deforming [97], and ultrasonic welding [98] methods, are also being investigated. These processes produce relatively low temperatures and are found to provide excellent joining without crystallization [99]. Among the supercooled-liquid welding techniques, friction is widely used for joining metallic-glass systems and has been studied by several groups [95,100]. The idea of joining metallic-glass systems using friction stir welding originated from the metal-extrusion technique. In metal extrusion, the strong bond within powder particles is established during the consolidation of the powder, due to superplastic deformation of the supercooled liquid [101]. Kamamura et al. [102] successfully joined Pd40Ni40P20 BMG using friction, taking advantage of the superplasticity of its supercooled liquid. The welded system was found to have properties similar to the parent system that was fabricated through the quenching technique. In addition, Wang et al. [103] studied the effect of friction time and rotational speed to maintain the glassy state within the interface and concluded that there is a critical time for any rotational speed to keep the interface below the crystallization temperature.
Using welding techniques to join disordered solids may allow the production of large metallic-glass systems, which can solve the issue with the requirement of a very high cooling rate. However, more research on this field will determine the intrinsic mechanism to join different metal-based MG systems. Different metallic-glass systems joined by liquid- and solid-state techniques are listed in Table 3.

3.3. Additive Manufacturing

In recent year, additive manufacturing (AM) has drawn much attention from the manufacturing industries and academia due its unique features that can significantly affect the microstructure and properties of engineering materials. This emerging additive manufacturing technology is also known as three-dimensional (3D) printing, in which a layer-by-layer fabrication process is used to build a 3D object directly from a CAD file [115]. Through additive manufacturing, it is possible to produce geometrically intricate designs, which would not be possible using conventional manufacturing techniques [115]. It is also possible to improve and customize material properties according to desired applications via the proper selection of process parameters during fabrication. Metallic-glass systems have also been produced by different additive manufacturing techniques [116,117,118,119]. Selective laser melting (SLM) is a branch of additive manufacturing, which is used for the formation of disordered solids [120,121,122] and does not require precise molds for the fabrication of complex and large-scale parts [123]. For example, SLM has been used to fabricate metallic-glass systems, which has been found to be beneficial as a composite matrix for biomedical applications [124,125].
In SLM, metal powder is placed on a metal plate and melted using a laser. The melt freezes rapidly and bonds with the metal plate. Additional layers of powder are then added as the melting and freezing steps are continued [123]. The overall process occurs in a closed chamber, where argon or helium is used to create an inert atmosphere to minimize contamination. The metal plate in the SLM process serves as the support for the final part and is usually iron (Fe), titanium (Ti), aluminum (Al) or nickel (Ni) [122,126] due to their outstanding mechanical stabilities and heat-dissipation characteristics [122]. Both the heating and cooling rates during the SLM process require the careful selection of parameters (laser power density, spot size, hatch spacing and scan speed) for the formation of a metallic glass [127]. The energy input into the molten powder is controlled by changing the spot size, the power of the laser, and the dwell time [128]. Optimizing these parameters can minimize the melt puddle, which is found to facilitate the glass formation [128].
The effects of scan speed and laser power to form a disordered structure have been studied by Jung et al. [129] for Fe-based (Fe68.3C6.9Si2.5B6.7P8.7Cr2.3Mo2.5Al2.1) metallic glass. Cross-section images of the resulting samples are shown in Figure 4. Jung’s study indicates that optimal energy transfer and high relative energy densities are helpful for forming metallic glass, which is achieved by using low scan speed and high laser power [129]. However, a study conducted by Li et al. indicates that using too high a laser energy density can introduce some crystallization into a system [130]. The impressive part of the layer-by-layer fabrication process of AM is that a refined microstructure and outstanding mechanical features can be achieved using SLM [129] as compared to samples produced via conventional fabrication techniques. However, crystallization is difficult to avoid due to the thermal processes involved in the SLM technique [127]. Moreover, thermal stress can induce cracks within the samples, which can effectively be mitigated by decreasing the energy density during the fabrication process [131]. Different metallic-glass systems fabricated using additive manufacturing process are listed in Table 4, along with some of the associated fabrication process parameters.

3.4. Powder Densification Technique

Metallic glasses produced by quenching techniques have size limitations (mm to cm) due to the requirement of very high cooling rates. A potential solution to avoid the issue of the cooling rate is the use of a fabrication route that consolidates amorphous powder. In this method, mechanical alloying [138,139] or high-pressure argon-gas atomization [140,141] is usually adopted to produce amorphous powders. Mechanical alloying is a simple, room-temperature method; however, it is a lengthy process. On the other hand, the high-pressure gas atomization technique requires an extreme experimental condition [142].
The amorphous powders are densified using other processes, such as extrusion methods [143,144,145], cold or hot pressing [146,147], spark-plasma sintering [139,148], and injection [149]. These densification processes are classified based on their operating temperature and time. For example, cold pressing [146] or equal channel angular extrusion [144] is performed at room temperature, whereas warm extrusion [150], hot pressing [147], injection [149], and spark-plasma sintering [142] require high temperatures. Among all these techniques, spark plasma sintering is popular due to densification capability within a short span of time [127]. Moreover, the spark plasma sintering process combined with dynamic pressing and fast heating is advantageous for the formation of a disordered solid, even for the marginal glass formers [127,151]. However, there is a possibility of crystallization if thermal stability is not controlled during the densification process [142].
Cardinal et al. [142,152] extensively studied a Cu-based metallic glass (Cu50Zr45Al5) fabricated via the spark plasma sintering method. In this study, planetary ball milling is used to make crystalline Zr (as shown in Figure 5a) and amorphous CuZrAl powders (Figure 5b). The obtained pure crystalline Zr (20% vol.) powder is homogeneously mixed with the amorphous CuZrAl powder (Figure 5c). The mixed powders (crystalline Zr and amorphous CuZrAl) are densified using the spark plasma sintering technique. As illustrated in Figure 6, the sintering temperature was 420 °C with an imposed pressure of 600 MPa. The application of high pressure and a fast-heating rate help to create a stronger bond between the powder grains and result in an amorphous solid [142].
The capability of porous metallic-glass systems to achieve outstanding mechanical properties (e.g., high strength, large elastic stains) and corrosion resistance has raised much research interest [153]. Conventional casting is found to hinder the successful formation of porous metallic glass structures due to the size constraints and the high contamination tendency [154]. In addition, spark plasma sintering alongside densification techniques are found to be a successful alternative to the methods that require very high cooling rates [155,156,157]. Different metallic-glass systems formed through the powder densification techniques are listed in Table 5.

3.5. Magnetron Co-Sputtering

Vapor quenching techniques, such as evaporation and magnetron co-sputtering, are a non-equilibrium vapor-to-solid transition method, which is often used to produce multicomponent metallic-glass systems [13]. In magnetron co-sputtering, the growth of the atoms on a substrate is governed by two competing factors, e.g., thermodynamics and kinetics [13]. When the energies of the arriving atoms were compared, the sputtered atoms were found to have average energy that is two orders of magnitude higher than that of the atoms in the evaporation technique [13]. Thus, from a thermodynamics perspective, the effective quench rate of the atoms is slower in sputtering than in evaporation [166]. However, from a kinetic point of view, the magnetron-sputtering process is specified in terms of momentum [13], and the targets with heavy atoms possess a higher momentum that facilitate the solidification of the atoms in a metallic glass system [167].
Sputtering systems are typically equipped with multiple guns, which are used to deposit materials on a substrate. Various fabrication parameters, such as the angle and power of each gun, gaseous environment, compositions of the targets, angle and rotation of the substrate, etc., can be carefully tuned to obtain optimized compositions of multicomponent metallic-glass systems [127]. By changing the aforementioned parameters, a compositional library of a metallic glass system can be synthesized for different engineering applications (as shown in Figure 7) [168]. Moreover, the slower cooling rate of sputtering and lack of nucleation sites of crystallization lead to properties that often emerge simultaneously in a single metallic-glass system. Bouala et al. [169] introduced Ag in Zr-Cu to form a compositional library of a Zr-Cu-Ag metallic glass system and studied the glass-forming ability, mechanical properties, electrochemical properties, and antibacterial activity of the fabricated systems for potential bio-medical applications. Among all the compositions, Zr73Cu16Ag11 metallic glass was found to exhibit optimum mechanical and electrochemical properties with enhanced antibacterial activity [169]. Metallic-glass systems produced by combinatorial development to achieve different properties are listed in Table 6.

4. Distinctive Properties

Like conventional crystalline materials, the properties of metallic-glass systems also originate from their microstructure, and the potential of metallic-glass systems as functional materials is an area of active research. The following sections summarize the structural, surface, mechanical, thermal, electrochemical, and magnetic characteristics of various metallic-glass systems that would be essential to solidify our understanding of these unique materials.

4.1. Structural and Surface Properties

Grain boundaries and secondary phase precipitates are the two common structural features that are absent in metallic-glass systems [176,177]. The homogeneous microstructure of metallic glass matrices contributes to their enhanced electrochemical and mechanical properties [35,177,178]. Furthermore, a better surface condition and minimal surface roughness are found to be useful in various engineering applications [58,179,180].
Various mathematical models have been proposed to explain the atomic structure of metallic-glass systems [181,182]. Among these models, Bernal’s dense random packing model is useful in describing the monoatomic amorphous structures of a short-range order [183]. However, the model is unable to explain the structures of binary metallic glasses [184]. In this regard, the model suggested by Gaskall [185] is rather useful but lacks proper experimental evidence. The model proposed by Miracle [186] considered a medium-range order (MRO) of metallic-glass systems with face-centered cubic (fcc) atomic packing. Other studies also considered a short- to medium-range order (MRO) for modeling different metallic-glass systems [187,188,189,190,191]. For example, the model suggested by Lee et al. [181] considered both short- and medium-range orders of binary Cu-Zr metallic glass. The short-range order of the Cu65Zr35 system exhibited a high packing density and a low free volume, whereas for the medium-range order, the majority of the icosahedra were found to exist in linked clusters. The icosahedra connected in the medium-range order resulted in the lowest average potential and atomic volume, which were found to provide more structural stability [192]. The structural modeling of metallic-glass systems postulates efficient packing around the solute and solvent atoms, and lower densities of the amorphous systems compared to their crystalline counterpart [193,194]. The hypothesis was later verified experimentally by Battezzati and Baricco, who reported lower molar volumes of different metallic-glass compositions [195]. Mukherjee et al. reported that the composed icosahedral structures connected in an extensive network with higher atomic packing led to a higher viscosity and lower atomic volume of metallic-glass systems [196]. Both of these two factors impact glass-forming ability and provide better structural properties of metallic-glass systems [197].
Different diffraction patterns and nano-scale characterization techniques have been carried out to analyze the disordered structure. The study conducted by Khan et al. [198] showed the microstructural analysis of a Zr-based (Zr-Ti-Fe-Al) metallic glass system using the HR-TEM technique. Figure 8 depicts the amorphous structure with embedded nano-crystallites, which was affirmed through selected-area electron diffraction (SAED) combined with fast Fourier transformation (FFT). The influence of inclusion on the properties of disordered solids was extensively examined by Liu et al. [199] by introducing different elements into the Cu-Ag system. It was found that the inclusion of Cr and Co in Cu-Ag improves electrical conductivity, with a slight compromise in strength, whereas the inclusion of Zr leads to a reverse trend. The relationship between the metallic glass composition, microstructure, and properties of various Zr-(Cu,Ag)-Al compositions was also reported by Jiang et al. and Chen et al. [62,197]. The synthesized metallic-glass systems are found to exhibit different physical and chemical properties with the change in sputtering power.
The composition of metallic glasses has been found to trigger different surface properties [197,199,200]. The increase in roughness with sputtering power indicates that under certain conditions, the power level can provide a composition with a better surface finish, whereas a higher deposition rate can be attributed to a higher surface roughness. Another study on Zr-based metallic glass (Zr–Ni–Al–Si) examined a broad range of compositions and nitrogen (0–17.7%) contents for improved physical and chemical features [61]. The inclusion of nitrogen of more than 7.1% was found to induce a higher level of crystallinity into the system, whereas the system without the inclusion of nitrogen was found to have a minimal amount of roughness. Moreover, thermally annealed metallic glass with an atomic composition of 60% Cu and 40% Zr exhibited an outstanding smooth surface, with fine grains and a dense microstructure [201]. In this study, thermal annealing was stated to impede the nucleation of nano-crystals due to the reduction in free volume in the atomic structure [201]. A study conducted by Ishii et al. also assessed the effect of thermal annealing in the atomic structure of metallic glass, where the structural analysis of the metallic glass system indicates the absence of pores with large open spaces, and it was predicted that the atomic rearrangement near the free volume does not occur due to the thermal treatment [202]. The study indicates that the metallic glasses may not consist of an absolute random structure; rather, the disordered state is influenced by the chemical order of the systems used for alloying [202]. Moreover, Zr-(Zr53Cu33Al9Ta5)- and Cu-(Cu48Zr42Ti4Al6)-based metallic glasses are found to exhibit a better surface condition, as the wettability of the surfaces indicates a significant reduction in the surface roughness from the crystalline materials [63]. A study conducted by Liu et al. [168] synthesized the compositional library for the Zr-based metallic-glass systems, and the disordered structural condition was found to exhibit lower roughness values compared to the crystalline materials.

4.2. Thermal Stability

The formation of metallic glass requires a proper understanding of thermodynamics, as the glass formation and observed material properties are interrelated. It is anticipated that the addition of specific elements in metallic-glass systems may often decrease the glass-forming ability [203]. A study conducted by Inoue et al. [204] indicated that the presence of Co, Cr, Fe, Mo, and V diminished the supercooled region due to a decrease in crystallization temperature (Tx), and, thus, reduced its glass-forming ability. The study determined that Hf inclusion in the Zr-based metallic glass kept the supercooled region unchanged, whereas the glass-phase area ratio extended with the inclusion of Ti, Nb, and Pd. Another study conducted by He et al. [203] analyzed the effect of Ta addition in the thermal stability of the supercooled region in a Zr-based system (Zr52.25Cu28.5Ni4.75Al9.5Ta5) fabricated using an arc-melting process. The study indicated that the addition of 3.2% Ta improved the thermal stability of the metallic glass matrix, although it did not significantly improve its glass-forming ability. For the Zr57Nb5Al10Cu15.4Ni12.6 metallic-glass system [205], the thermal stability of the system was tested by adding Mo, Nb, and Ta. It was observed that the addition of these species did not alter its thermal stability. Inoue and Zhang studied the thermal stability of Cu-Zr-Al metallic-glass systems [206] for a wide range of compositions, and it was found that the inclusion of Al (3–10%) increased the temperature difference (ΔT) between the glass-transition temperature (Tg) and the crystallization temperature (Tx), which further stabilized the systems thermodynamically. For an Fe-based Fe43Cr16Mo16C15B10 metallic glass [207], negative heat-of-mixing values were found within the range of 1–45 kJ mol−1, which is within the range of stabilization for a supercooled liquid.
The empirical rules of Inoue [48] associated with the negative enthalpy between the atomic pairs are also related to the thermal stability of metallic-glass systems [207]. It has been observed that the systems that follow this particular rule usually form a denser and long-range configuration (by suppressing the nucleation for crystallization) with a relatively lower atomic diffusivity and a higher viscosity [207,208]. For example, the formation of Fe-based metallic glass (Fe43Cr16Mo16C15B10) with a higher Δ T is attributed to the low atomic diffusivity, nucleation, and growth reactions among the species [207]. Apart from the multicomponent systems, the glass-forming ability and thermal behavior of binary systems were also studied. For example, Zr-Cu systems of an extensive compositional range [50] were characterized using differential scanning calorimetry. The systems sputtered at a low-density discharge condition exhibited an optimum range of the stable supercooled region with the inclusion of 31–62 at% Cu, whereas the high-density discharge condition resulted in an inconsistent supercooled region. Finally, the thermal stability of the supercooled region, the atomic species, and the fabrication routes all govern the electrochemical [207], mechanical [203,205] and structural [202] responses of metallic-glass systems. Among different fabrication routes, the combinatorial method via sputtering has been reported to yield thermodynamically stable metallic-glass systems and has been adopted for easy glass formation [50,64,175,209].

4.3. Mechanical Properties

When metallic-glass systems are considered for structural, biomedical, defense, and aerospace applications, the evaluation of mechanical properties is essential [35,36,45,210,211]. A higher yield strength and increased elastic limit are the two properties that make metallic-glass systems distinct from commonly used crystalline materials [35,69]. Metallic-glass systems have been reported to have significantly higher yield strength and hardness values compared to 316 L stainless steel-, Ti-, and Zr-based alloys [35,75,212,213]. Moreover, a lower elastic modulus and comparatively higher elastic strain have also been reported for different metallic-glass systems, which are beneficial for applications where a reduction in stress concentration is vital [214,215,216,217]. The yield strength is the ability of a material to resist plastic deformation, and metallic-glass systems are found to have a higher yield strength [218]. Metallic-glass systems also exhibit higher wear resistance compared to crystalline materials. Moreover, the longer interatomic distance and other unique microstructural features often help to obtain a lower elastic modulus and outstanding fracture strength [218,219,220,221,222]. Different characterization techniques, such as nanoindentation, the micro-hardness test, and compressive and tensile tests, are performed to extract the properties of metallic-glass systems. The mechanical properties of various Zr-, Ti-, Fe-, Mg-, Cu-, Pd-, Al-, Ni-, Ca-based systems have been listed in Table 7.
Disordered-state solids have the limitations of a reduced plasticity at room temperature. Metallic glass systems are reported to deform catastrophically under localized shear bands due to their limited number of active shear bands [69,251,252,253,254]. The combination of the strength and ductility of disordered solids has been reported to have improved with the introduction of nano-crystallites [255,256] or nano-quasicrystals [256,257]. Calin et al. [69] suggested that the inclusion of nano-crystalline may be helpful to improve their ductility. A study conducted by Liu et al. [258] established a relationship between the strength and ductility of CuZr-based metallic-glass composites, as shown in Figure 9a. The study indicated that the combination of high strength and improved ductility can be achieved through percolation by matching the length scales of particle size and inter-particle distance in the amorphous microstructure [258]. The relationship between the tensile strength and ductility among different metallic-glass systems is shown in Figure 9b. Furthermore, presence of specific species in a multicomponent system is another crucial factor to consider in achieving strength and ductility in an alloy. Eckert et al. [259] fabricated a binary metallic glass (Cu50Zr50) with nano-crystallites and compared its response with a ternary metallic-glass system (Cu47.5Zr47.5Al5). It was reported that the inclusion of Al in the matrix extended the nucleation of its shear bands [259]. In this study, the multiple homogenous nucleation of shear bands in the ternary metallic glass system (Cu47.5Zr47.5Al5) was found to exhibit a better compressive ductility alongside an increased strength [259]. However, to provide a specific hypothesis on how to achieve strength and ductility in a metallic-glass system requires rigorous study and is currently a research challenge.

4.4. Electrochemical Properties

Homogeneity and the absence of a long-range order in the microstructure have been found to be advantageous to better electrochemical properties [264,265]. The dissolution rate of a multi-component system depends on the alloying element species, chemical composition, electrolyte chemistry, experimental environment and thermodynamic metastability [35,266,267,268]. The effects of compositional and chemical homogeneity, as well as the effects of a short- to medium-range order, on the electrochemical responses of disordered solids have been studied by different research groups [29,269]. A study conducted by Kou et al. [270] found a better corrosion resistance in Zr- and Ti-based metallic glasses (Zr41.25Ti13.75Ni10Cu12.5Be22.5 and Ti40Zr25Ni8Cu9Be18) in NaCl and H2SO4 solutions compared to a crystalline Zr41.25Ti13.75Ni10Cu12.5Be22.5 alloy. It was concluded that chemical homogeneity and the absence of crystalline defects facilitate the passive behavior of metallic-glass systems compared to the crystalline counterpart [270]. Furthermore, metallic glasses were found to be thermodynamically metastable; thus, their structures tend to exhibit the excessive release of free volume [35,271]. A study conducted by Jiang et al. [271] determined the electrochemical behavior of Zr52.5Cu17.9Ni14.6Al10.0Ti5.0 metallic glass due to the change in free volume. The study indicated that the reduction in free volume improves the corrosion resistance of metallic-glass systems [271].
Metallic-glass systems were also found to have a higher resistance towards localized or pitting corrosion compared to crystalline materials [220,272]. One of the primary reasons behind pitting corrosion is the presence of physical irregularities in the protective oxide or passive layer [266,273]. Different phases, scratches, grain boundaries, and crystal imperfections can cause damage in the passive layer and eventually trigger catastrophic failure of the systems [274]. The absence of long-range-order and corrosion initiation sites in the lattice structure were found to facilitate the electrochemical properties of metallic-glass systems [275,276].
An assessment of the electrochemical properties of different metallic glasses has been conducted in a simulated environment using different electrolytes for various engineering applications [177,277]. The commonly used electrolytes are phosphate buffer solution (PBS) [266,278,279], Hank’s solution [154,229,280], sodium sulfate (Na2SO4) [58,281], Ringer’s solution [275,282], aqueous NaCl solution [209,283], etc. Among the aforementioned electrolytes, PBS (pH = 7.4), Hank’s (pH = 7.4) and Ringer’s (pH = 6.5) solutions are commonly used to explore the electrochemical behavior of metallic-glass systems for bio-implant applications, whereas 0.05 M Na2SO4 and 0.6 M NaCl mimic humid air and simulate seawater environments, respectively [154,177,229,278]. Among various glass-forming materials, Zr-, Ti-, and Fe- based metallic-glass systems have been extensively studied [272,284,285,286,287,288,289,290]. The enthusiasm behind the selection of these metallic-glass systems has been attributed to their higher glass-forming abilities [20,291,292], lower ion dissolution rates, and stronger and faster oxide-forming capabilities within a wide pH range [293,294,295]. In addition, the presence of oxide-forming and noble species can also influence the rate of corrosion. The presence of Nb [296], Ag [169], Cu [297], Co [298], Cr [290,299], Ni [300], Mo [277], Be [301], Al [302], W [283], and Pd [229] in metallic-glass systems have been found to improve their electrochemical properties. A study conducted by Zhang et al. [303] explored the effect of Ti, Cr, Mn, Fe, Co, Ni and Cu inclusion in an Al-based metallic-glass system. Its polarization curves (as shown in Figure 10) indicate excellent corrosion resistance, a higher pitting potential, and a comparable or lower corrosion current density in the metallic-glass systems compared to pure Al [303]. The study concluded that engineered metallic-glass systems with quasi- or nano-crystalline phases are electrochemically superior to conventional crystalline materials [303].

4.5. Magnetic Properties

Excellent soft magnetic properties make metallic-glass systems ideal candidate materials for sensors, actuators, and magneto-optic storage applications [304,305,306]. The magnetic properties of metallic-glass systems were found to be improved by thermal annealing [307]. The significance of thermal annealing is the formation of nano-crystallites within the amorphous system [307]; thus, the amorphous–crystalline composite microstructure shows enhanced softness compared to a completely disordered microstructure [308]. The size and distribution of the nano-crystals also influence the overall magnetic properties [309]. A mathematical model by Jin et al. [309] explained that the grain orientation changes the coercivity and remanence, and thus, modifies the magnetic properties of a system.
Fe- [310,311] and Co-based [312] metallic-glass systems are commonly investigated due to their outstanding magnetic features, such as their low coercivity, high electrical resistivity, and saturation magnetization [304]. These inherent magnetic properties are also dependent on the chemical composition and atomic configuration of the species of the amorphous systems [129]. A study conducted by Zhu et al. [313] showed a relationship between the coercivity and saturation magnetization of Fe-based metallic-glass systems due to the inclusion of various amounts of Mo (as shown in Figure 11). A higher saturation magnetization is attributed to the modification of the local moment due to the inclusion of Mo. In addition, the elimination of pinning spots creating domain walls was found to be useful in achieving a lower coercivity, which yields soft magnetic properties to the system [313]. Furthermore, Shen et al. [314] explored the magnetic properties of Fe-based Fe–Ga–P–C–B–Si metallic-glass systems in detail. The study indicated that Fe-based metallic glasses exhibit an improved Currie temperature, saturated magnetization, and coercive force [314], which are indicative of good soft magnetic properties. Another study conducted by Lin et al. [315] considered the influence of Yttrium (Y), Dysprosium (Dy), Holmium (Ho), Erbium (Er), and Scandium (Sc) inclusion in an Fe-B-based alloy. The study reported that the species with 130% atomic size mismatched with Fe, and a eutectic point with both Fe and B may impart high saturation magnetization and electrical resistivity, and low coercivity to the amorphous Fe-based solids [315]. Furthermore, a Co-based metallic-glass system (Co–Fe–Ta–B) studied by Inoue et al. [316] exhibited low coercive force (0.25 A/m) and high permeability (550,000). The promising soft magnetic properties of the Co-based metallic glass were attributed to the homogenous atomic structure of the system [317]. The magnetic properties of different Fe- and Co-based metallic-glass systems are listed in Table 8.

5. Applications of Metallic Glasses

5.1. Biomedical Applications

Metallic glasses can be a perfect alternative to conventional crystalline biomaterials (such as 316L stainless steel, Ti or Ti-based alloys, Zr or Zr-based alloys, Co-Cr alloys, etc.) when used as coatings for surgical devices and implants inside the human body. The following two sections (Section 5.1.1 and Section 5.1.2) highlight some applications of metallic glasses in biomedical sectors.

5.1.1. Antibacterial Application

Nosocomial infections are often escalated due to bacterial infections from medical instruments or devices. Commonly known biomaterials have been found to be ineffective at preventing bacterial infections [58,336,337,338], whereas metallic-glass systems exhibit excellent antibacterial properties. This unique characteristic is attributed to the composition of their multicomponent amorphous systems, their lower surface roughness, and the presence of antibacterial species in the matrices [63,168,265]. Liu et al. studied the antibacterial capabilities of Zr-Cu-Al-Ag systems, and the observed antibacterial responses are shown in Figure 12 [168]. The Zr38Cu36Al18Ag8 system, which had a lower surface roughness, was found to have the highest antibacterial activity in this study [168]. Hydrophobic surfaces are well known for their better antibacterial response, and metallic-glass systems have been found to exhibit better wettability compared to crystalline materials due to their disordered microstructure [339]. The presence of certain species, such as silver (Ag) and copper (Cu), in a multicomponent system provides excellent resistance towards bacterial attack [59,169,340]. For example, a Zr39Cu39Ag22 metallic glass was found to be very efficient against S. aureus [58], whereas Zr61Al7.5Ni10Cu17.5Si4 was found to be efficient against S. aureus, E. coli, A. baumannii, P. aeruginosa, and C. albicans bacteria [341]. Another study conducted by Chu et al. [63] compared the antibacterial activity of Zr-based (Zr53Cu33Al9Ta5) and Cu-based (Cu48Zr42Ti4Al6) metallic-glass systems. The result showed that the adhesions of E. coli and S. aureus bacteria were greatly hindered on metallic glasses than those on a bare Si wafer [63]. For E. coli, both metallic-glass systems showed a 100% antibacterial activity, whereas the Zr-based metallic glass was more antibacterial towards the S. aureus than the E. Coli bacteria [63].
Commercially used surgical blades are usually made of stainless steel that contains micron-scale roughness on the edge tips and surfaces. The roughness hinders the smooth cut of the soft tissues, and the resulting wear and tear are often difficult to recover. Such limitations of surgical blades and scissors can be resolved by using metallic-glass systems, which exhibit exceptional surface characteristics [179]. A study conducted by Tsai et al. [179] reported lower roughness values, a higher blade sharpness index, and a lower depth of indentation when using a Zr-based metallic-glass system (Zr48Cu35.3Al8Ag8Si0.7). All of these characteristics make metallic-glass systems an optimum solution for surgical instruments without a compromise in their performance [179].

5.1.2. Bio-Implants

Durability and biosafety are two characteristics that are extremely desirable for implantable materials. Elastic modulus mismatch between an implant and a bone, and a lack of resistance towards localized corrosion have hindered the growth of conventional crystalline materials (such as 316L stainless steel, Co-Cr alloys, and Ti-based alloys) as premium choices for implantable materials. The lower corrosion resistance of crystalline materials is often attributed to the presence of grain boundaries and phase precipitates, which act as preferred regions for adverse electrochemical reactions [35,54,336,342,343]. Furthermore, the short interatomic distances of crystalline materials result in higher elastic modulus values, which may lead to failure of the implants due to a stress shielding effect [218]. Metallic-glass systems have the potential to mitigate this problem due to their homogenous microstructures, with longer interatomic distances and the absence of grain boundaries [35].
Due to their excellent electrochemical and mechanical properties, metallic-glass systems are being used as vascular stents, and dental and orthopedic implants. Zr- [212,273,273,284,344,345], Fe- [275,290,346], and Ti- based [347,348,349] metallic-glass systems have been widely explored in in vitro conditions for permanent implants. The biological responses of pre-osteoblast cells (MC3T3-E1) [180,350,351,352,353], fibroblast cells (L929 and NIH3T3) [302,344,346,354], human-osteoblast-like cells (SaOS2 and MG63) [348,355,356], and endothelial cells [297] reveal the outstanding bio-compatible characteristics of different multicomponent metallic-glass systems. Qiu et al. reported excellent mechanical (improved strength and plasticity), electrochemical (a low passive current density and high pitting potential), and biocompatible responses of a Zr-based system (Zr60Cu22.5Pd5Al7.5Nb5) in an embryonic-mouse-fibroblast cell line (NIH3T3 cell) [285]. The better biocompatibility of the metallic-glass system was attributed to the short/medium-range order and oxide-forming capability of the amorphous structure [285]. Fe-based metallic-glass systems, studied by Li et al. [290], exhibited better biocompatibility towards NIH3T3 cells and better electrochemical responses in artificial saliva compared to conventional biomaterials. Furthermore, Zr-based metallic glasses have been considered for cardiovascular stents, and studied for endothelial and muscle cells [297]. A cell-morphology and cell-metabolic-activity assessment, as shown in Figure 13, revealed the faster growth of endothelial (HAECs) cells on the Zr-based metallic glasses than on 316 L stainless steel, whereas the growth of smooth muscle (HASCMs) cells was relatively slower [297]. That study reported higher endothelial cell-adhesion capabilities on the Zr-based metallic glasses compared to their conventional crystalline counterpart [297].
Research on biodegradable metallic glass is also gaining a lot of enthusiasm. The biocompatibility, mechanical properties, and electrochemical responses of Mg- [357,358,359,360,361], Ca- [362,363], Sr- [364], and Zn-based [365] degradable alloys have been studied extensively by several research groups. However, researchers investigating biodegradable metallic-glass systems used as fully functional bio-implants still face significant research challenges in their obtaining optimum mechanical and electrochemical properties. For example, bio-implants typically require a higher strength; however, degradation due to pitting corrosion creates surface defects that lead to a gradual loss of strength [35]. In addition, there is a possibility of tissue damage due to hydrogen evolution under a body-fluid environment [35]. Therefore, a proper understanding of this degradation mechanism and the relationship between its strength and degradation rate are required.

5.2. Electrochemical Devices

The lower efficiency and durability of catalysts are two primary obstacles that hinder the growth of electrochemical devices to meet the rising energy demand. Owing to their outstanding electrocatalytic activity and durability, metallic glasses can be considered prominent candidates for energy-storage and -conversion devices, such as fuel and electrolysis cells, and batteries. Nevertheless, some key components of electrochemical devices, such as the membrane, catalyst, and separator, still require further development to resolve issues associated with weight and cost [366]. For example, proton-exchange membrane (PEM) fuel cells are suitable candidates for power-generation applications [367,368,369]. PEM fuel cells consist of bipolar plates that are electrically conductive and essential to isolating the fuel and oxidant gases. Bipolar plates are conventionally made of carbon graphite; however, its brittleness and high production cost limit its use. To overcome this limitation, Kim et al. [367] investigated Ni65Cr15P16B4-metallic-glass-coated plates fabricated using HVOF spray coating and a subsequent hot-pressing technique. Enhanced corrosion resistance and durability suggest the potential of a multicomponent metallic-glass system to be used as bipolar plates in fuel cells [367]. Another prerequisite of a fuel cell is efficiency, and ultra-high-purity hydrogen was found to be crucial to enhancing its efficiency [370]. Crystalline materials are currently being used for hydrogen-storage applications [371]. But the embrittlement of crystalline materials limits its applications. A novel multicomponent metallic glass could be an alternate choice, as the glassy matrix of metallic glass possesses numerous sites for hydrogen absorption [371] and requires no modification to its microstructure [372]. A study conducted by Jayalakshmi et al. [370] explored Ni-Nb metallic-glass systems for hydrogen-related energy applications. The study determined the higher absorption-capacity and embrittlement-resistance characteristics of metallic-glass systems, which induce the better interaction of hydrogen with the metals [370]. Moreover, high permeability and a lower dissolution of metallic glasses make them ideal choices for hydrogen-permeable membranes and separators of fuel cells, respectively [370].

5.3. Optoelectronic Devices

Metallic-glass systems are gaining substantial momentum in the micro- and nano-imprinting of optoelectronic devices [373]. Smooth surface conditions and the negative enthalpy of mixing of metallic-glass systems facilitate optical transmittance and reflectivity, which are essential for different optoelectronic devices [304]. In the optoelectronic industry, indium-tin-oxide (ITO) is a prominent choice due to its excellent transparency and conductivity [374]. ITO films are also often used in solar cells and collectors, automobile windows, camera lenses, and lamps [374,375]. However, the cost of using ITO films is very high and the use of a metallic glass structure can be a prominent alternative [304]. Huang et al. studied a bi-layer ITO/ZrCu structure, which achieved good conductivity and transparency [376]. The study discovered that the lower resistivity and negative enthalpy between the atoms of the metallic-glass system can be beneficial by forming a continuous layer for transparent conductor design [376]. Wang et al. [377] investigated a Ag40Mg18Al42 metallic glass system and found that the lower surface roughness, atomic defects, free volumes, and electric resistivity improved its optical reflectivity. It is important to note that some of these features, such as surface morphology, atomic structure, and chemical composition, depend on fabrication routes and can be improved further by post-heat-treatment processes [376,378].
Chalcogenide glasses have found numerous applications in optical devices as well. These unique glasses are based on chalcogen elements such as S, Se, and Te, and are formed by the addition of Ge, As, Sb, and Ga and doped by rare-earth elements. A review article by Seddon [379] discusses their fabrication in bulk, fiber, and film form, their optical and thermal properties, and their applications.

5.4. Aerospace Application

The use of disordered-state solids in aerospace applications is gaining interest, although no concrete study has been conducted on any multi-component system until recently. Their high strength and lightweight attributes extend the demand of metallic-glass systems into lighter, smaller, and cost-effective aerospace applications [210]. Research conducted by Axinte [210] indicated that the aircraft and spacecraft fasteners can be produced from metallic glass. Another study by Burgess et al. [380] indicated that the higher strength and hardness of metallic-glass systems may be useful for coatings in aerospace applications. Aluminum alloys (Al-6061 and Al-7075) are widely used in automotive and aerospace industries due to their lighter weight and enhanced thermal conductivity [381]. Telford studied the possibility of using Al in a metallic-glass combination [211]. However, its limited glass-forming ability (GFA), higher affinity of oxide formation, and the need for an extreme experimental condition imposed a research challenge on the development of Al-based metallic-glass systems [381]. Vitreloy-1 (Zr41.2Ti13.8Cu12.5Ni10.0Be22.5) [11] is the only commercially available metallic glass that is being studied by the US Department of Energy and NASA for aerospace applications [211]. The primary limitation of using metallic-glass systems in industrial applications is their low plasticity, i.e., minimal plastic deformation before catastrophic failure. However, the combination of higher strength and better corrosion resistance may offset this limitation to foster industrial applications.

5.5. Memory Storage Devices

Non-volatile memory (NVM) is being used in computers, smartphones, flash memory devices, and other electronic devices. However, the requirement for a higher density and voltage hinders the writing capacity of NVM [382]. To find a suitable alternative, metal-oxide thin films are being used as storage devices or resistive random-access memory [383,384,385]. Tulu et al. investigated and optimized a thin-film metallic-glass oxide (TFMGO) for resistive switching and as a multicomponent oxide memory device [386]. A 15 nm thick oxide film of (ZrCuAlNi)Ox was fabricated on a Pt/Ti/Si substrate using magnetron sputtering in the presence of oxygen. The RS I-V curve, as shown in Figure 14a shows the current–voltage formation of a Pt/TFMGO/Pt device with a unipolar behavior of switching. Moreover, its endurance features, as shown in Figure 14b, indicate no reduction in resistance up to 1250 switching cycles [386]. The outstanding features are further confirmed by its retention characteristics, as shown in Figure 14c, which exhibit a good resistance ratio and the regeneration of resistive switching without the need of thermal forming [386]. These excellent behaviors are attributed to the amorphosity of the metallic-glass matrix, which has been confirmed through nano-scale characterization. These findings create an opportunity to use thin-sized metallic-glass systems for storage device applications.

6. Conclusions

This article presents a comprehensive review of the history, current state, and prospect of various metallic-glass systems for various engineering applications. The formation of metallic glasses has been discussed in light of the ‘confusion principle’ of Greer and the empirical rules of Inoue. As postulated by Inoue, three different parameters, i.e., the number of atomic species, atomic size ratio, and negative heat of mixing affect glass formation, and have been discussed in detail; however, metallic-glass formation often deviates from these empirical rules. Different conventional and novel fabrication routes, their prospects, and limitations have also been highlighted. The requirement of a very high critical cooling rate of conventional arc-melting or die casting hinders the commercial production of disordered structures. The issues of the critical cooling rate and limited glass-forming ability can be overcome through combinatorial development. However, the limitation of size restricts the potential industrial applications of metallic-glass systems. The ongoing research is focused on mitigating the size constraints of metallic-glass systems via different welding techniques. Nonetheless, metallic-glass systems have been termed the next-generation materials due to their outstanding mechanical, electrochemical, thermal, magnetic, and optical properties. These properties are attributed to the random atomic arrangements of amorphous solids compared to conventional crystalline materials. Furthermore, the absence of dislocations and grain boundaries in a metallic-glass matrix provides better corrosion resistance in different aggressive mediums. Different features of metallic-glass systems, their limitations, and their applications are summarized in Figure 15. However, limited plasticity and the possibility of catastrophic failure restrict metallic-glass high-load-bearing structural applications. Therefore, there are ample research opportunities to improve the mechanical properties of amorphous solids. For example, adding minor elements such as; Al and La on the matrix of the multicomponent systems has been observed to enhance the ductility and plasticity of high-entropy alloys [387]. Recently, metallic glass systems by adding La or Ce is fabricated successfully [388], however, the effects of such addition on the physical properties requires further investigation. Similarly, composite structures embedded with nano-particles are also well known for excellent mechanical and microstructural properties [389]. Hence, such use of nano-particles in metallic-glass structures can also be explored. Furthermore, metallic glasses in thin film forms have been gaining popularity due to their unique mechanical and electrochemical properties and make such structures suitable for surgical devices, implants, medical devices, supercapacitors, and battery applications [265,273]. However, the toxicity of certain glass-forming elements poses challenges on their use in biological environments. A higher-saturation magnetization and Currie temperature have also been reported for the metallic-glass systems, which enables them to be used as magnetic materials for different applications. In addition, metallic-glass systems have been found to have wide super-cooled regions with a high glass-transition temperature. These outstanding features make metallic-glass systems distinct alternatives to crystalline materials for different engineering applications, including biological implants, surgical tools, electrochemical, optoelectronic, and memory storage applications.

Author Contributions

Conceptualization, A.J., I.S. and W.H.; methodology, A.J. and M.N.B.; formal analysis, A.J., M.N.B., W.H. and I.S.; investigation, A.J.; data curation, A.J. and M.N.B.; writing—original draft preparation, A.J.; writing—review and editing, A.J., M.N.B., W.H. and I.S.; visualization, A.J.; supervision, W.H. and I.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

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. Chen, M. A brief overview of bulk metallic glasses. NPG Asia Mater. 2011, 3, 82–90. [Google Scholar] [CrossRef]
  2. Klement, W.; Willens, R.H.; Duwez, P. Non-crystalline Structure in Solidified Gold–Silicon Alloys. Nature 1960, 187, 869–870. [Google Scholar] [CrossRef]
  3. Chen, H.S.; Turnbull, D. Formation, stability and structure of palladium-silicon based alloy glasses. Acta Metall. 1969, 17, 1021–1031. [Google Scholar] [CrossRef]
  4. Chen, H.S. Thermodynamic considerations on the formation and stability of metallic glasses. Acta Metall. 1974, 22, 1505–1511. [Google Scholar] [CrossRef]
  5. Inoue, A. Stabilization of metallic supercooled liquid and bulk amorphous alloys. Acta Mater. 2000, 48, 279–306. [Google Scholar] [CrossRef]
  6. 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]
  7. Drehman, A.J.; Greer, A.L.; Turnbull, D. Bulk formation of a metallic glass: Pd40Ni40P20. Appl. Phys. Lett. 1982, 41, 716–717. [Google Scholar] [CrossRef]
  8. Kui, H.W.; Greer, A.L.; Turnbull, D. Formation of bulk metallic glass by fluxing. Appl. Phys. Lett. 1984, 45, 615–616. [Google Scholar] [CrossRef]
  9. Inoue, A.; Zhang, T.; Masumoto, T. Zr–Al–Ni Amorphous Alloys with High Glass Transition Temperature and Significant Supercooled Liquid Region. Mater. Trans. JIM 1990, 31, 177–183. [Google Scholar] [CrossRef]
  10. Schroers, J. Processing of bulk metallic glass. Adv. Mater. 2010, 22, 1566–1597. [Google Scholar] [CrossRef]
  11. 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]
  12. Inoue, A.; Takeuchi, A. Recent development and application products of bulk glassy alloys. Acta Mater. 2011, 59, 2243–2267. [Google Scholar] [CrossRef]
  13. Braeckman, B.R.; Depla, D. On the amorphous nature of sputtered thin film alloys. Acta Mater. 2016, 109, 323–329. [Google Scholar] [CrossRef]
  14. Yokoyama, Y.; Mund, E.; Inoue, A.; Schultz, L. Production of Zr55Cu30Ni5Al10 Glassy Alloy Rod of 30 mm in Diameter by a Cap-Cast Technique. Mater. Trans. 2007, 48, 3190–3192. [Google Scholar] [CrossRef]
  15. Inoue, A.; Nishiyama, N.; Kimura, H. Preparation and Thermal Stability of Bulk Amorphous Pd40Cu30Ni10P20 Alloy Cylinder of 72 mm in Diameter. Mater. Trans. JIM 1997, 38, 179–183. [Google Scholar] [CrossRef]
  16. Zeng, Y.; Nishiyama, N.; Yamamoto, T.; Inoue, A. Ni-Rich Bulk Metallic Glasses with High Glass-Forming Ability and Good Metallic Properties. Mater. Trans. 2009, 50, 2441–2445. [Google Scholar] [CrossRef]
  17. Zheng, Q.; Xu, J.; Ma, E. High glass-forming ability correlated with fragility of Mg–Cu(Ag)–Gd alloys. J. Appl. Phys. 2007, 102, 113519. [Google Scholar] [CrossRef]
  18. Zhang, Q.; Zhang, W.; Inoue, A. Preparation of Cu36Zr48Ag8Al8 Bulk Metallic Glass with a Diameter of 25 mm by Copper Mold Casting. Mater. Trans. 2007, 48, 629–631. [Google Scholar] [CrossRef]
  19. Ponnambalam, V.; Poon, S.J.; Shiflet, G.J. Fe-based bulk metallic glasses with diameter thickness larger than one centimeter. J. Mater. Res. 2004, 19, 1320–1323. [Google Scholar] [CrossRef]
  20. Zhu, S.L.; Wang, X.M.; Inoue, A. Glass-forming ability and mechanical properties of Ti-based bulk glassy alloys with large diameters of up to 1 cm. Intermetallics 2008, 16, 1031–1035. [Google Scholar] [CrossRef]
  21. Dina, S.; Geyer, U.; Minnigerode, G.V. Investigations on macroscopic intrinsic stress in amorphous binary-alloy films. Ann. Phys. 1992, 504, 164–171. [Google Scholar] [CrossRef]
  22. Geyer, U.; von Hülsen, U.; Kopf, H. Internal interfaces and intrinsic stress in thin amorphous Cu-Ti and Co-Tb films. J. Appl. Phys. 1998, 83, 3065–3070. [Google Scholar] [CrossRef]
  23. Thomas, R.E.; Perepezko, J.H.; Wiley, J.D. Crystallization of sputter deposited amorphous metal thin films. Appl. Surf. Sci. 1986, 26, 534–541. [Google Scholar] [CrossRef]
  24. Qing-Ming, C.; Yu-Dian, F.; Heng-De, L. Amorphization of binary alloys by magnetron cosputtering. Mater. Lett. 1988, 6, 311–315. [Google Scholar] [CrossRef]
  25. Chen, H.S. Glassy metals. Rep. Prog. Phys. 1980, 43, 353–432. [Google Scholar] [CrossRef]
  26. Zhang, T.; Inoue, A. Density, Thermal Stability and Mechanical Properties of Zr–Ti–Al–Cu–Ni Bulk Amorphous Alloys with High Al Plus Ti Concentrations. Mater. Trans. JIM 1998, 39, 857–862. [Google Scholar] [CrossRef]
  27. Schuh, C.A.; Hufnagel, T.C.; Ramamurty, U. Mechanical behavior of amorphous alloys. Acta Mater. 2007, 55, 4067–4109. [Google Scholar] [CrossRef]
  28. Byrne, C.J.; Eldrup, M. Bulk Metallic Glasses. Science 2008, 321, 502–503. [Google Scholar] [CrossRef]
  29. Archer, M.D.; Corke, C.C.; Harji, B.H. The electrochemical properties of metallic glasses. Electrochim. Acta 1987, 32, 13–26. [Google Scholar] [CrossRef]
  30. Carmo, M.; Sekol, R.C.; Ding, S.; Kumar, G.; Schroers, J.; Taylor, A.D. Bulk Metallic Glass Nanowire Architecture for Electrochemical Applications. ACS Nano 2011, 5, 2979–2983. [Google Scholar] [CrossRef]
  31. O’Handley, R.C.; Hasegawa, R.; Ray, R.; Chou, C.-P. Ferromagnetic properties of some new metallic glasses. Appl. Phys. Lett. 1976, 29, 330–332. [Google Scholar] [CrossRef]
  32. Hasegawa, R.; O’Handley, R.C. Soft magnetic properties of metallic glasses—Recent developments. J. Appl. Phys. 1979, 50, 1551–1556. [Google Scholar] [CrossRef]
  33. Jabed, A.; Meduri, C.S.; Kumar, G. Effect of time on the isothermal viscosity of metallic glass supercooled liquids. J. Alloys Compd. 2021, 863, 158067. [Google Scholar] [CrossRef]
  34. Jagdale, S.; Jabed, A.; Theeda, S.; Meduri, C.S.; Hu, Z.; Hasan, M.; Kumar, G. Review of Thermoplastic Drawing with Bulk Metallic Glasses. Metals 2022, 12, 518. [Google Scholar] [CrossRef]
  35. Li, H.F.; Zheng, Y.F. Recent advances in bulk metallic glasses for biomedical applications. Acta Biomater. 2016, 36, 1–20. [Google Scholar] [CrossRef]
  36. Schroers, J.; Kumar, G.; Hodges, T.M.; Chan, S.; Kyriakides, T.R. Bulk metallic glasses for biomedical applications. JOM 2009, 61, 21–29. [Google Scholar] [CrossRef]
  37. Hata, S.; Sato, K.; Shimokohbe, A. Fabrication of thin film metallic glass and its application to microactuators. Proc. SPIE 1999, 3892, 97–108. [Google Scholar] [CrossRef]
  38. Liu, Y.; Hata, S.; Wada, K.; Shimokohbe, A. Thermal, Mechanical and Electrical Properties of Pd-Based Thin-Film Metallic Glass. Jpn. J. Appl. Phys. 2001, 40, 5382–5388. [Google Scholar] [CrossRef]
  39. 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]
  40. Luber, E.; Mohammadi, R.; Ophus, C.; Lee, Z.; Nelson-Fitzpatrick, N.; Westra, K.; Evoy, S.; Dehman, U.; Radmilovic, V.; Mitlin, D. Tailoring the microstructure and surface morphology of metal thin films for nano-electro-mechanical systems applications. Nanotechnology 2008, 19, 125705. [Google Scholar] [CrossRef]
  41. Hu, Y.C.; Wang, Y.Z.; Su, R.; Cao, C.R.; Li, F.; Sun, C.W.; Yang, Y.; Guan, P.F.; Ding, D.W.; Wang, Z.L.; et al. A Highly Efficient and Self-Stabilizing Metallic-Glass Catalyst for Electrochemical Hydrogen Generation. Adv. Mater. 2016, 28, 10293–10297. [Google Scholar] [CrossRef] [PubMed]
  42. Baiker, A. Metallic glasses in heterogeneous catalysis. Faraday Discuss. Chem. Soc. 1989, 87, 239. [Google Scholar] [CrossRef]
  43. Katona, T.; Molnar, A. Amorphous Alloy Catalysis: VII. Activation and Surface Characterization of an Amorphous Cu-Ti Alloy Catalyst Precursor in the Dehydrogenation of 2-Propanol and Comparison with Cu-Zr1. J. Catal. 1995, 153, 333–343. [Google Scholar] [CrossRef]
  44. Kruzic, J.J. Bulk Metallic Glasses as Structural Materials: A Review. Adv. Eng. Mater. 2016, 18, 1308–1331. [Google Scholar] [CrossRef]
  45. Ashby, M.F.; Greer, A.L. Metallic glasses as structural materials. Scr. Mater. 2006, 54, 321–326. [Google Scholar] [CrossRef]
  46. Greer, A.L. Confusion by design. Nature 1993, 366, 303–304. [Google Scholar] [CrossRef]
  47. Perim, E.; Lee, D.; Liu, Y.; Toher, C.; Gong, P.; Li, Y.; Simmons, W.N.; Levy, O.; Vlassak, J.J.; Schroers, J.; et al. Spectral descriptors for bulk metallic glasses based on the thermodynamics of competing crystalline phases. Nat. Commun. 2016, 7, 12315. [Google Scholar] [CrossRef]
  48. Takeuchi, A.; Inoue, A. Classification of bulk metallic glasses by atomic size difference, heat of mixing and period of constituent elements and its application to characterization of the main alloying element. Mater. Trans. 2005, 46, 2817–2829. [Google Scholar] [CrossRef]
  49. Li, Y.; Zhao, S.; Liu, Y.; Gong, P.; Schroers, J. How Many Bulk Metallic Glasses Are There? ACS Comb. Sci. 2017, 19, 687–693. [Google Scholar] [CrossRef]
  50. Zeman, P.; Zítek, M.; Zuzjaková, S.; Čerstvý, R. Amorphous Zr-Cu thin-film alloys with metallic glass behavior. J. Alloys Compd. 2017, 696, 1298–1306. [Google Scholar] [CrossRef]
  51. Wang, W.H. Bulk Metallic Glasses with Functional Physical Properties. Adv. Mater. 2009, 21, 4524–4544. [Google Scholar] [CrossRef]
  52. He, Y.; Poon, S.J.; Shiflet, G.J. Synthesis and properties of metallic glasses that contain aluminum. Science 1988, 241, 1640–1642. [Google Scholar] [CrossRef] [PubMed]
  53. Johnson, W.L. Fundamental Aspects of Bulk Metallic Glass Formation in Multicomponent Alloys. Mater. Sci. Forum 1996, 225–227, 35–50. [Google Scholar] [CrossRef]
  54. Calin, M.; Gebert, A.; Ghinea, A.C.; Gostin, P.F.; Abdi, S.; Mickel, C.; Eckert, J.H. Designing biocompatible Ti-based metallic glasses for implant applications. Mater. Sci. Eng. C 2013, 33, 875–883. [Google Scholar] [CrossRef]
  55. Schroers, J.; Johnson, W.L. Ductile Bulk Metallic Glass. Phys. Rev. Lett. 2004, 93, 255506. [Google Scholar] [CrossRef]
  56. Hofmann, D.C.; Suh, J.-Y.; Wiest, A.; Duan, W.; Lind, M.-L.; Demetriou, M.D.; Johnson, W.L. Designing metallic glass matrix composites with high toughness and tensile ductility. Nature 2008, 451, 1085–1089. [Google Scholar] [CrossRef]
  57. Das, J.; Tang, M.B.; Kim, K.B.; Theissmann, R.; Baier, F.; Wang, W.H.; Eckert, J.H. Work-Hardenable’ Ductile Bulk Metallic Glass. Phys. Rev. Lett. 2005, 94, 205501. [Google Scholar] [CrossRef]
  58. Etiemble, A.; Loughian, C.D.; Apreutesei, M.; Langlois, C.; Cardinal, S.; Pelletier, J.M.; Pierson, J.-F.; Steyer, P. Innovative Zr-Cu-Ag thin film metallic glass deposed by magnetron PVD sputtering for antibacterial applications. J. Alloys Compd. 2017, 707, 155–161. [Google Scholar] [CrossRef]
  59. Chu, Y.Y.Y.; Lin, Y.S.S.; Chang, C.M.M.; Liu, J.-K.K.; Chen, C.H.H.; Huang, J.C.C. Promising antimicrobial capability of thin film metallic glasses. Mater. Sci. Eng. C 2014, 36, 221–225. [Google Scholar] [CrossRef]
  60. Lou, B.S.; Yang, Y.C.; Lee, J.W.; Chen, L.T. Biocompatibility and mechanical property evaluation of Zr-Ti-Fe based ternary thin film metallic glasses. Surf. Coat. Technol. 2017, 320, 512–519. [Google Scholar] [CrossRef]
  61. Chuang, C.-Y.; Lee, J.-W.; Li, C.-L.; Chu, J.P. Mechanical properties study of a magnetron-sputtered Zr-based thin film metallic glass. Surf. Coat. Technol. 2013, 215, 312–321. [Google Scholar] [CrossRef]
  62. Chen, H.W.; Hsu, K.-C.; Chan, Y.-C.; Duh, J.-G.; Lee, J.-W.; Ching Jang, J.S.; Chen, G.-J. Antimicrobial properties of Zr-Cu-Al-Ag thin film metallic glass. Thin Solid Film. 2014, 561, 98–101. [Google Scholar] [CrossRef]
  63. Chu, J.P.; Liu, T.-Y.; Li, C.-L.; Wang, C.-H.; Jang, J.S.C.; Chen, M.-J.; Chang, S.-H.; Huang, W.-C. Fabrication and characterizations of thin film metallic glasses: Antibacterial property and durability study for medical application. Thin Solid Film. 2014, 561, 102–107. [Google Scholar] [CrossRef]
  64. Cai, C.-N.N.; Zhang, C.; Sun, Y.-S.S.; Huang, H.-H.H.; Yang, C.; Liu, L. ZrCuFeAlAg thin film metallic glass for potential dental applications. Intermetallics 2017, 86, 80–87. [Google Scholar] [CrossRef]
  65. Zhang, Z.F.; Eckert, J.; Schultz, L. Difference in compressive and tensile fracture mechanisms of Zr59Cu20Al10Ni8Ti3 bulk metallic glass. Acta Mater. 2003, 51, 1167–1179. [Google Scholar] [CrossRef]
  66. Yao, K.F.; Zhang, C.Q. Fe-based bulk metallic glass with high plasticity. Appl. Phys. Lett. 2007, 90, 061901. [Google Scholar] [CrossRef]
  67. Kobayashi, A.; Yano, S.; Kimura, H.; Inoue, A. Fe-based metallic glass coatings produced by smart plasma spraying process. Mater. Sci. Eng. B 2008, 148, 110–113. [Google Scholar] [CrossRef]
  68. Kim, Y.C.; Kim, W.T.; Kim, D.H. A development of Ti-based bulk metallic glass. Mater. Sci. Eng. A 2004, 375–377, 127–135. [Google Scholar] [CrossRef]
  69. Calin, M.; Eckert, J.; Schultz, L. Improved mechanical behavior of Cu-Ti-based bulk metallic glass by in situ formation of nanoscale precipitates. Scr. Mater. 2003, 48, 653–658. [Google Scholar] [CrossRef]
  70. Liu, Z.; Chan, K.C.; Liu, L. Enhanced glass forming ability and plasticity of a Ni-free Zr-based bulk metallic glass. J. Alloys Compd. 2009, 487, 152–156. [Google Scholar] [CrossRef]
  71. Kawashima, A.; Wada, T.; Ohmura, K.; Xie, G.; Inoue, A. A Ni- and Cu-free Zr-based bulk metallic glass with excellent resistance to stress corrosion cracking in simulated body fluids. Mater. Sci. Eng. A 2012, 542, 140–146. [Google Scholar] [CrossRef]
  72. Greer, A.L. Effect of quench rate on the structural relaxation of a metallic glass. J. Mater. Sci. 1982, 17, 1117–1124. [Google Scholar] [CrossRef]
  73. Inoue, A.; Kato, A.; Zhang, T.; Kim, S.G.; Masumoto, T. Mg Cu–Y Amorphous Alloys with High Mechanical Strengths Produced by a Metallic Mold Casting Method. Mater. Trans. JIM 1991, 32, 609–616. [Google Scholar] [CrossRef]
  74. Greer, A.L. Metallic Glasses. Science 1995, 267, 1947–1953. [Google Scholar] [CrossRef] [PubMed]
  75. Lin, C.H.; Huang, C.H.; Chuang, J.F.; Lee, H.C.; Liu, M.C.; Du, X.H.; Huang, J.C.; Jang, J.S.C.; Chen, C.H. Simulated body-fluid tests and electrochemical investigations on biocompatibility of metallic glasses. Mater. Sci. Eng. C 2012, 32, 2578–2582. [Google Scholar] [CrossRef]
  76. Srivastava, A.P.; Babu, D.A.; Verma, A.; Deshmukh, A.A.; Kaushal, A.; Palikundwar, U.A. Understanding the effect of Hf on thermal stability and glass forming ability of Fe57.2Co30.8Zr7-xHfxB4Cu1 (x = 3, 5, and 7) metallic glasses. J. Non-Cryst. Solids 2019, 503–504, 7–12. [Google Scholar] [CrossRef]
  77. Nagel, C.; Rätzke, K.; Schmidtke, E.; Wolff, J.; Geyer, U. Free-volume changes in the bulk metallic glass and the undercooled liquid. Phys. Rev. B Condens. Matter Mater. Phys. 1998, 57, 10224–10227. [Google Scholar] [CrossRef]
  78. Davies, H.A.; Aucote, J.; Hull, J.B. Amorphous Nickel produced by Splat Quenching. Nat. Phys. Sci. 1973, 246, 13–14. [Google Scholar] [CrossRef]
  79. Davies, H.A. The kinetics of formation of A AuGeSi metallic glass. J. Non-Cryst. Solids 1975, 17, 266–272. [Google Scholar] [CrossRef]
  80. Woychik, C.G.; Lowndes, D.H.; Massalski, T.B. Solidification structures in melt-spun and pulsed laser-quenched Cu-Ti alloys. Acta Metall. 1985, 33, 1861–1871. [Google Scholar] [CrossRef]
  81. Massalski, T.B.; Woychik, C.G.; Dutkiewicz, J. Solidification structures in rapidly quenched Cu-Ti-Zr alloys. Metall. Trans. A 1988, 19, 1853–1860. [Google Scholar] [CrossRef]
  82. Lin, C.; Spaepen, F. Fe--B glasses formed by picosecond pulsed laser quenching. Appl. Phys. Lett. 1982, 41, 721–723. [Google Scholar] [CrossRef]
  83. Lin, C.-J.; Spaepen, F. Nickel-niobium alloys obtained by picosecond pulsed laser quenching. Acta Metall. 1986, 34, 1367–1375. [Google Scholar] [CrossRef]
  84. McCluskey, P.J.; Vlassak, J.J. Combinatorial nanocalorimetry. J. Mater. Res. 2010, 25, 2086–2100. [Google Scholar] [CrossRef]
  85. Gregoire, J.M.; McCluskey, P.J.; Dale, D.; Ding, S.; Schroers, J.; Vlassak, J.J. Combining combinatorial nanocalorimetry and X-ray diffraction techniques to study the effects of composition and quench rate on Au–Cu–Si metallic glasses. Scr. Mater. 2012, 66, 178–181. [Google Scholar] [CrossRef]
  86. Kuball, A.; Stolpe, M.; Busch, R. Crystallization behavior of the Al86Ni8Y6 metallic glass forming alloy upon rapid cooling. J. Alloys Compd. 2018, 737, 398–404. [Google Scholar] [CrossRef]
  87. Zhong, L.; Wang, J.; Sheng, H.; Zhang, Z.; Mao, S.X. Formation of monatomic metallic glasses through ultrafast liquid quenching. Nature 2014, 512, 177–180. [Google Scholar] [CrossRef]
  88. Zhu, Y.Y.; Li, Z.G.; Li, R.F.; Li, M.; Daze, X.L.; Feng, K.; Wu, Y.X. Microstructure and property of Fe–Co–B–Si–C–Nb amorphous composite coating fabricated by laser cladding process. Appl. Surf. Sci. 2013, 280, 50–54. [Google Scholar] [CrossRef]
  89. Kawamura, Y.; Ohno, Y. Successful Electron-Beam Welding of Bulk Metallic Glass. Mater. Trans. 2001, 42, 2476–2478. [Google Scholar] [CrossRef]
  90. Li, B.; Li, Z.Y.; Xiong, J.G.; Xing, L.; Wang, D.; Li, Y. Laser welding of Zr45Cu48Al7 bulk glassy alloy. J. Alloys Compd. 2006, 413, 118–121. [Google Scholar] [CrossRef]
  91. Zhou, Y.; Zhang, Q.; He, G.; Guo, J. Connection of bulk amorphous alloy Zr55Al10Ni5Cu30 by high current density electropulsing. Mater. Lett. 2003, 57, 2208–2211. [Google Scholar] [CrossRef]
  92. Kim, J.; Shin, S.; Lee, C. Characterization of the Gas Tungsten Arc Welded Cu54Ni6Zr22Ti18 Bulk Metallic Glass Weld. Mater. Trans. 2005, 46, 1440–1442. [Google Scholar] [CrossRef]
  93. Pakdil, M.; Çam, G.; Koçak, M.; Erim, S. Microstructural and mechanical characterization of laser beam welded AA6056 Al-alloy. Mater. Sci. Eng. A 2011, 528, 7350–7356. [Google Scholar] [CrossRef]
  94. Wang, G.; Huang, Y.J.; Shagiev, M.; Shen, J. Laser welding of Ti40Zr25Ni3Cu12Be20 bulk metallic glass. Mater. Sci. Eng. A 2012, 541, 33–37. [Google Scholar] [CrossRef]
  95. Kawamura, Y.; Shoji, T.; Ohno, Y. Welding technologies of bulk metallic glasses. J. Non-Cryst. Solids 2003, 317, 152–157. [Google Scholar] [CrossRef]
  96. Kawamura, Y.; Ohno, Y.; Chiba, A. Development of Welding Technologies in Bulk Metallic Glasses. Mater. Sci. Forum 2002, 13, 553–558. [Google Scholar] [CrossRef]
  97. Chen, W.; Liu, Z.; Schroers, J. Joining of bulk metallic glasses in air. Acta Mater. 2014, 62, 49–57. [Google Scholar] [CrossRef]
  98. Kim, J. Weldability of Cu54Zr22Ti18Ni6 bulk metallic glass by ultrasonic welding processing. Mater. Lett. 2014, 130, 160–163. [Google Scholar] [CrossRef]
  99. Sun, Y.; Ji, Y.; Fujii, H.; Nakata, K.; Nogi, K. Microstructure and mechanical properties of friction stir welded joint of Zr55Cu30Al10Ni5 bulk metallic glass with pure copper. Mater. Sci. Eng. A 2010, 527, 3427–3432. [Google Scholar] [CrossRef]
  100. Shin, H.-S.; Jeong, Y.-J.; Choi, H.-Y.; Kato, H.; Inoue, A. Friction Welding of Zr55Al10Ni5Cu30 Bulk Metallic Glasses. Mater. Trans. 2005, 46, 2768–2772. [Google Scholar] [CrossRef]
  101. ACE Powder Metallurgy Ltd. Metal Powder Report No. 52; Elsevier Advanced Technology: Amsterdam, The Netherlands, 1946; Volume 10. [Google Scholar]
  102. Kawamura, Y.; Ohno, Y. Superplastic bonding of bulk metallic glasses using friction. Scr. Mater. 2001, 45, 279–285. [Google Scholar] [CrossRef]
  103. Wang, G.; Huang, Y.J.; Makhanlall, D.; Shen, J. Friction joining of Ti40Zr25Ni3Cu12Be20 bulk metallic glass. J. Mater. Process. Technol. 2012, 212, 1850–1855. [Google Scholar] [CrossRef]
  104. Kagao, S.; Kawamura, Y.; Ohno, Y. Electron-beam welding of Zr-based bulk metallic glasses. Mater. Sci. Eng. A 2004, 375–377, 312–316. [Google Scholar] [CrossRef]
  105. Kawamura, Y.; Kagao, S.; Ohno, Y. Electron Beam Welding of Zr-Based Bulk Metallic Glass to Crystalline Zr Metal. Mater. Trans. 2001, 42, 2649–2651. [Google Scholar] [CrossRef]
  106. Kim, J.; Kawamura, Y. Electron beam welding of the dissimilar Zr-based bulk metallic glass and Ti metal. Scr. Mater. 2007, 56, 709–712. [Google Scholar] [CrossRef]
  107. Kim, J.; Kawamura, Y. Dissimilar welding of Zr41Be23Ti14Cu12Ni10 bulk metallic glass and stainless steel. Scr. Mater. 2011, 65, 1033–1036. [Google Scholar] [CrossRef]
  108. Kim, J.H.; Lee, C.; Lee, D.M.; Sun, J.H.; Shin, S.Y.; Bae, J.C. Pulsed Nd:YAG laser welding of Cu54Ni6Zr22Ti18 bulk metallic glass. Mater. Sci. Eng. A 2007, 449–451, 872–875. [Google Scholar] [CrossRef]
  109. Kawahito, Y.; Terajima, T.; Kimura, H.; Kuroda, T.; Nakata, K.; Katayama, S.; Inoue, A. High-power fiber laser welding and its application to metallic glass Zr55Al10Ni5Cu30. Mater. Sci. Eng. B 2008, 148, 105–109. [Google Scholar] [CrossRef]
  110. Wang, H.-S.; Chen, H.-G.; Jang, J.S.-C. Microstructure evolution in Nd:YAG laser-welded (Zr53Cu30Ni9Al8)Si0.5 bulk metallic glass alloy. J. Alloys Compd. 2010, 495, 224–228. [Google Scholar] [CrossRef]
  111. Shao, L.; Datye, A.; Huang, J.; Ketkaew, J.; Sohn, S.W.; Zhao, S.; Wu, S.; Zhang, Y.; Schwarz, U.D.; Schorers, J. Pulsed Laser Beam Welding of Pd43Cu27Ni10P20 Bulk Metallic Glass. Sci. Rep. 2017, 7, 7989. [Google Scholar] [CrossRef]
  112. Shin, H.-S.; Jung, Y.-C. Characteristics of friction stir spot welding of Zr-based bulk metallic glass sheets. J. Alloys Compd. 2010, 504, S279–S282. [Google Scholar] [CrossRef]
  113. Shin, H.-S.; Jeong, Y.-J.; Choi, H.-Y.; Kato, H.; Inoue, A. Joining of Zr-based bulk metallic glasses using the friction welding method. J. Alloys Compd. 2007, 434–435, 102–105. [Google Scholar] [CrossRef]
  114. Ohkubo, T.; Shoji, S.; Kawamura, Y.; Hono, K. Nanostructure analysis of friction welded Pd–Ni–P/Pd–Cu–Ni–P metallic glass interface. Scr. Mater. 2005, 53, 493–497. [Google Scholar] [CrossRef]
  115. Al-Mamun, N.S.; Deen, K.M.; Haider, W.; Asselin, E.; Shabib, I. Corrosion behavior and biocompatibility of additively manufactured 316L stainless steel in a physiological environment: The effect of citrate ions. Addit. Manuf. 2020, 34, 101237. [Google Scholar] [CrossRef]
  116. Li, Y.; Shen, Y.; Leu, M.C.; Tsai, H.-L. Mechanical properties of Zr-based bulk metallic glass parts fabricated by laser-foil-printing additive manufacturing. Mater. Sci. Eng. A 2019, 743, 404–411. [Google Scholar] [CrossRef]
  117. Lu, Y.; Huang, Y.; Wu, J. Laser additive manufacturing of structural-graded bulk metallic glass. J. Alloys Compd. 2018, 766, 506–510. [Google Scholar] [CrossRef]
  118. Wu, W.; Jiang, J.; Li, G.; Fuh, J.Y.H.; Jiang, H.; Gau, 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]
  119. Mahbooba, Z.; Thorsson, L.; Unosson, M.; Skoglund, P.; West, H.; Horn, T.; Rock, C.; Vogli, E.; Harrysson, O. Additive manufacturing of an iron-based bulk metallic glass larger than the critical casting thickness. Appl. Mater. Today 2018, 11, 264–269. [Google Scholar] [CrossRef]
  120. Ouyang, D.; Li, N.; Xing, W.; Zhang, J.; Liu, L. 3D printing of crack-free high strength Zr-based bulk metallic glass composite by selective laser melting. Intermetallics 2017, 90, 128–134. [Google Scholar] [CrossRef]
  121. Ouyang, D.; Li, N.; Liu, L. Structural heterogeneity in 3D printed Zr-based bulk metallic glass by selective laser melting. J. Alloys Compd. 2018, 740, 603–609. [Google Scholar] [CrossRef]
  122. Santos, E.C.; Shiomi, M.; Osakada, K.; Laoui, T. Rapid manufacturing of metal components by laser forming. Int. J. Mach. Tools Manuf. 2006, 46, 1459–1468. [Google Scholar] [CrossRef]
  123. Pauly, S.; Löber, L.; Petters, R.; Stoica, M.; Scudino, S.; Kühn, U.; Eckert, J.H. Processing metallic glasses by selective laser melting. Mater. Today 2013, 16, 37–41. [Google Scholar] [CrossRef]
  124. Chen, P.S.; Tsai, P.H.; Li, T.H.; Jang, J.S.; Huang, J.C.; Lin, C.H.; Pan, C.T.; Lin, H.K. Development and Fabrication of Biocompatible Ti-Based Bulk Metallic Glass Matrix Composites for Additive Manufacturing. Materials 2023, 16, 5935. [Google Scholar] [CrossRef] [PubMed]
  125. Aliyu, A.A.; Panwisawas, C.; Shinjo, J.; Puncreobutr, C.; Reed, R.C.; Poungsiri, K.; Lohwongwatana, B. Laser-based additive manufacturing of bulk metallic glasses: Recent advances and future perspectives for biomedical applications. J. Mater. Res. Technol. 2023, 23, 2956–2990. [Google Scholar] [CrossRef]
  126. Elahinia, M.H.; Hashemi, M.; Tabesh, M.; Bhaduri, S.B. Manufacturing and processing of NiTi implants: A review. Prog. Mater. Sci. 2012, 57, 911–946. [Google Scholar] [CrossRef]
  127. 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]
  128. Zheng, B.; Zhou, Y.; Smugeresky, J.E.; Lavernia, E.J. Processing and behavior of Fe-based metallic glass components via laser-engineered net shaping. Int. J. Powder Metall. 2009, 45, 27–39. [Google Scholar] [CrossRef]
  129. Jung, H.Y.; Choi, S.J.; Prashanth, K.G.; Stoica, M.; Scudino, S.; Yi, S.; Kühn, U.; Kim, D.H.; Kim, K.B.; Eckert, J.H. Fabrication of Fe-based bulk metallic glass by selective laser melting: A parameter study. Mater. Des. 2015, 86, 703–708. [Google Scholar] [CrossRef]
  130. Li, X.P.; Roberts, M.P.; O’Keeffe, S.; Sercombe, T.B. Selective laser melting of Zr-based bulk metallic glasses: Processing, microstructure and mechanical properties. Mater. Des. 2016, 112, 217–226. [Google Scholar] [CrossRef]
  131. Li, X.P.; Kang, C.W.; Huang, H.; Sercombe, T.B. The role of a low-energy–density re-scan in fabricating crack-free Al85Ni5Y6Co2Fe2 bulk metallic glass composites via selective laser melting. Mater. Des. 2014, 63, 407–411. [Google Scholar] [CrossRef]
  132. Li, X.P.; Roberts, M.; Liu, Y.J.; Kang, C.W.; Huang, H.; Sercombe, T.B. Effect of substrate temperature on the interface bond between support and substrate during selective laser melting of Al–Ni–Y–Co–La metallic glass. Mater. Des. 2015, 65, 1–6. [Google Scholar] [CrossRef]
  133. Deng, L.; Wang, S.; Wang, P.; Kühn, U.; Pauly, S. Selective laser melting of a Ti-based bulk metallic glass. Mater. Lett. 2018, 212, 346–349. [Google Scholar] [CrossRef]
  134. Pauly, S.; Schricker, C.; Scudino, S.; Deng, L.; Kühn, U. Processing a glass-forming Zr-based alloy by selective laser melting. Mater. Des. 2017, 135, 133–141. [Google Scholar] [CrossRef]
  135. Li, N.; Zhang, J.; Xing, W.; Ouyang, D.; Liu, L. 3D printing of Fe-based bulk metallic glass composites with combined high strength and fracture toughness. Mater. Des. 2018, 143, 285–296. [Google Scholar] [CrossRef]
  136. Nong, X.D.; Zhou, X.L.; Ren, Y.X. Fabrication and characterization of Fe-based metallic glasses by Selective Laser Melting. Opt. Laser Technol. 2019, 109, 20–26. [Google Scholar] [CrossRef]
  137. Lu, Y.; Zhang, H.; Li, H.; Xu, H.; Huang, G.; Qin, Z.; Lu, X. Crystallization prediction on laser three-dimensional printing of Zr-based bulk metallic glass. J. Non-Cryst. Solids 2017, 461, 12–17. [Google Scholar] [CrossRef]
  138. Zheng, B.; Ashford, D.; Zhou, Y.; Mathaudhu, S.N.; Delplanque, J.-P.; Lavernia, E.J. Influence of mechanically milled powder and high pressure on spark plasma sintering of Mg–Cu–Gd metallic glasses. Acta Mater. 2013, 61, 4414–4428. [Google Scholar] [CrossRef]
  139. Li, X.; Cheng, Z.; Hu, K.; Chen, H.; Yang, C. Crystallization kinetics and spark plasma sintering of amorphous Ni53Nb20Ti10Zr8Co6Ta3 powders prepared by mechanical alloying. Vacuum 2015, 114, 93–100. [Google Scholar] [CrossRef]
  140. Xie, G.Q.; Louzguine-Luzgin, D.V.; Fukuhara, M.; Kimura, H.; Inoue, A. Consolidation Behavior of Cu-Zr-Al Metallic Glass Powder by Spark Plasma Sintering. Mater. Sci. Forum 2010, 654–656, 1086–1089. [Google Scholar] [CrossRef]
  141. Kim, C.K.; Lee, S.; Shin, S.Y.; Kim, D.H. Effects of consolidation temperature and pressure on microstructures and mechanical properties of Cu-based bulk amorphous alloys consolidated by spark plasma sintering. J. Alloys Compd. 2008, 453, 108–114. [Google Scholar] [CrossRef]
  142. Cardinal, S.; Pelletier, J.M.; Qiao, J.C.; Bonnefont, G.; Xie, G. Influence of spark plasma sintering parameters on the mechanical properties of Cu50Zr45Al5 bulk metallic glass obtained using metallic glass powder. Mater. Sci. Eng. A 2016, 677, 116–124. [Google Scholar] [CrossRef]
  143. Lee, M.H.; Bae, D.H.; Kim, W.T.; Kim, D.H.; Rozhkova, E.; Wheelock, P.B.; Sordelet, D.J. Synthesis of Ni-based bulk amorphous alloys by warm extrusion of amorphous powders. J. Non-Cryst. Solids 2003, 315, 89–96. [Google Scholar] [CrossRef]
  144. Robertson, J.; Im, J.-T.; Karaman, I.; Hartwig, K.T.; Anderson, I.E. Consolidation of amorphous copper based powder by equal channel angular extrusion. J. Non-Cryst. Solids 2003, 317, 144–151. [Google Scholar] [CrossRef]
  145. Sordelet, D.J.; Rozhkova, E.; Huang, P.; Wheelock, P.B.; Besser, M.F.; Kramer, M.J.; Calvo-Dahlborg, M.; Dahlborg, U. Synthesis of Cu47Ti34Zr11Ni8 Bulk Metallic Glass by Warm Extrusion of Gas Atomized Powders. J. Mater. Res. 2002, 17, 186–198. [Google Scholar] [CrossRef]
  146. Kim, Y.B.; Jang, D.H.; Seok, H.K.; Kim, K.Y. Fabrication of Fe–Si–B based amorphous powder cores by cold pressing and their magnetic properties. Mater. Sci. Eng. A 2007, 449–451, 389–393. [Google Scholar] [CrossRef]
  147. Kim, Y.B.; Park, H.M.; Jeung, W.Y.; Bae, J.S. Vacuum hot pressing of gas-atomized Ni–Zr–Ti–Si–Sn amorphous powder. Mater. Sci. Eng. A 2004, 368, 318–322. [Google Scholar] [CrossRef]
  148. Deng, S.S.; Wang, D.J.; Luo, Q.; Huang, Y.J.; Shen, J. Spark plasma sintering of gas atomized AlNiYLaCo amorphous powders. Adv. Powder Technol. 2015, 26, 1696–1701. [Google Scholar] [CrossRef]
  149. Kim, C.K.; Son, C.-Y.; Ha, D.J.; Yoon, T.S.; Lee, S.; Kim, N.J. Microstructure and mechanical properties of powder-injection-molded products of Cu-based amorphous powders and Fe-based metamorphic powders. Mater. Sci. Eng. A 2008, 476, 69–77. [Google Scholar] [CrossRef]
  150. Cai, A.; Xiong, X.; Liu, Y.; An, W.; Zhou, G.; Lou, Y.; Li, T.; Li, X. Effect of consolidation parameters on mechanical properties of Cu-based bulk amorphous alloy consolidated by hot pressing. Trans. Nonferrous Met. Soc. China 2012, 22, 2032–2040. [Google Scholar] [CrossRef]
  151. Kelly, J.P.; Fuller, S.M.; Seo, K.; Novitskaya, E.; Eliasson, V.; Hodge, A.M.; Graeve, O.A. Designing in situ and ex situ bulk metallic glass composites via spark plasma sintering in the super cooled liquid state. Mater. Des. 2016, 93, 26–38. [Google Scholar] [CrossRef]
  152. Cardinal, S.; Pelletier, J.M.; Xie, G.Q.; Mercier, F. Manufacturing of Cu-based metallic glasses matrix composites by spark plasma sintering. Mater. Sci. Eng. A 2018, 711, 405–414. [Google Scholar] [CrossRef]
  153. Lee, M.H.; Sordelet, D.J. Nanoporous metallic glass with high surface area. Scr. Mater. 2006, 55, 947–950. [Google Scholar] [CrossRef]
  154. Xie, G.; Qin, F.; Zhu, S.; Louzguine-Lugzin, D.V. Corrosion behaviour of porous Ni-free Ti-based bulk metallic glass produced by spark plasma sintering in Hanks’ solution. Intermetallics 2014, 44, 55–59. [Google Scholar] [CrossRef]
  155. Xie, G.; Zhang, W.; Louzguine-Luzgin, D.V.; Kimura, H.; Inoue, A. Fabrication of porous Zr–Cu–Al–Ni bulk metallic glass by spark plasma sintering process. Scr. Mater. 2006, 55, 687–690. [Google Scholar] [CrossRef]
  156. Lee, M.H.; Sordelet, D.J. Synthesis of bulk metallic glass foam by powder extrusion with a fugitive second phase. Appl. Phys. Lett. 2006, 89, 021921. [Google Scholar] [CrossRef]
  157. Qiu, K.; Yu, B.; Ren, Y. Porous bulk metallic glass fabricated by powder hot pressing. J. Univ. Sci. Technol. Beijing Miner. Metall. Mater. 2007, 14, 59–63. [Google Scholar] [CrossRef]
  158. Choi, P.P.; Kim, J.S.; Nguyen, O.T.H.; Kwon, D.H.; Kwon, Y.S.; Kim, J.C. Al-La-Ni-Fe bulk metallic glasses produced by mechanical alloying and spark-plasma sintering. Mater. Sci. Eng. A 2007, 449–451, 1119–1122. [Google Scholar] [CrossRef]
  159. Xie, G.; Louzguine-Luzgin, D.V.; Kimura, H.; Inoue, A. Nearly full density Ni52.5Nb10Zr15Ti15Pt7.5 bulk metallic glass obtained by spark plasma sintering of gas atomized powders. Appl. Phys. Lett. 2007, 90, 241902. [Google Scholar] [CrossRef]
  160. Lee, J.K.; Kim, H.J.; Kim, T.S.; Shin, S.Y.; Kim, Y.C.; Bae, J.C. Deformation behavior of Ni-based bulk metallic glass synthesized by spark plasma sintering. J. Mater. Process. Technol. 2007, 187–188, 801–804. [Google Scholar] [CrossRef]
  161. Lee, S.; Kato, H.; Kubota, T.; Makino, A.; Inoue, A. Fabrication and soft-magnetic properties of Fe–B–Nb–Y glassy powder compacts by spark plasma sintering technique. Intermetallics 2009, 17, 218–221. [Google Scholar] [CrossRef]
  162. Xie, G.; Louzguine-Luzgin, D.V.; Song, L.; Kimura, H.; Inoue, A. Dual phase metallic glassy composites with large-size and ultra-high strength fabricated by spark plasma sintering. Intermetallics 2009, 17, 512–516. [Google Scholar] [CrossRef]
  163. Harimkar, S.P.; Paital, S.R.; Singh, A.; Aalund, R.; Dahotre, N.B. Microstructure and properties of spark plasma sintered Fe–Cr–Mo–Y–B–C bulk metallic glass. J. Non-Cryst. Solids 2009, 355, 2179–2182. [Google Scholar] [CrossRef]
  164. Ishihara, S.; Zhang, W.; Kimura, H.; Omori, M.; Inoue, A. Consolidation of Fe-Co-Nd-Dy-B Glassy Powders by Spark-Plasma Sintering and Magnetic Properties of the Consolidated Alloys. Mater. Trans. 2003, 44, 138–143. [Google Scholar] [CrossRef]
  165. Li, Q.; Wang, G.; Song, X.; Fan, L.; Hu, W.; Xiao, F.; Yang, Q.; Ma, M.; Zhang, J.; Liu, R. Ti50Cu23Ni20Sn7 bulk metallic glasses prepared by mechanical alloying and spark-plasma sintering. J. Mater. Process. Technol. 2009, 209, 3285–3288. [Google Scholar] [CrossRef]
  166. Mahieu, S.; Depla, D. Reactive sputter deposition of TiN layers: Modelling the growth by characterization of particle fluxes towards the substrate. J. Phys. D Appl. Phys. 2009, 42, 053002. [Google Scholar] [CrossRef]
  167. Van Steenberge, S.; Leroy, W.P.; Hubin, A.; Depla, D. Momentum transfer driven textural changes of CeO2 thin films. Appl. Phys. Lett. 2014, 105, 111602. [Google Scholar] [CrossRef]
  168. Liu, Y.; Padmanabhan, J.; Cheung, B.; Liu, J.; Chen, Z.; Scanley, B.E.; Wesolowski, D.; Pressley, M.; Broadbridge, C.C.; Altman, S.; et al. Combinatorial development of antibacterial Zr-Cu-Al-Ag thin film metallic glasses. Sci. Rep. 2016, 6, 26950. [Google Scholar] [CrossRef] [PubMed]
  169. Bouala, G.I.N.; Etiemble, A.; Der Loughian, C.; Langlois, C.; Pierson, J.-F.; Steyer, P. Silver influence on the antibacterial activity of multi-functional Zr-Cu based thin film metallic glasses. Surf. Coat. Technol. 2018, 343, 108–114. [Google Scholar] [CrossRef]
  170. Chu, C.W.; Jang, J.S.C.; Chen, G.J.; Chiu, S.M. Characteristic studies on the Zr-based metallic glass thin film fabricated by magnetron sputtering process. Surf. Coat. Technol. 2008, 202, 5564–5566. [Google Scholar] [CrossRef]
  171. Deng, Y.P.; Guan, Y.F.; Fowlkes, J.D.; Wen, S.Q.; Liu, F.X.; Pharr, G.M.; Liaw, P.K.; Liu, C.T.; Rack, P.D. A combinatorial thin film sputtering approach for synthesizing and characterizing ternary ZrCuAl metallic glasses. Intermetallics 2007, 15, 1208–1216. [Google Scholar] [CrossRef]
  172. Chou, H.S.; Huang, J.C.; Chang, L.W. Mechanical properties of ZrCuTi thin film metallic glass with high content of immiscible tantalum. Surf. Coat. Technol. 2010, 205, 587–590. [Google Scholar] [CrossRef]
  173. Ketov, S.V.; Shi, X.; Xie, G.; Kumashiro, R.; Churyumov, A.Y.; Bazlov, A.I.; Chen, N.; Ishikawa, Y.; Asao, N.; Wu, H.; et al. Nanostructured Zr-Pd metallic glass thin film for biochemical applications. Sci. Rep. 2015, 5, 7799. [Google Scholar] [CrossRef] [PubMed]
  174. Liu, F.X.; Yang, F.Q.; Gao, Y.F.; Jiang, W.h.; Guan, Y.F.; Rack, P.D.; Sergic, O.; Liaw, P.K. Micro-scratch study of a magnetron-sputtered Zr-based metallic-glass film. Surf. Coat. Technol. 2009, 203, 3480–3484. [Google Scholar] [CrossRef]
  175. Lai, J.J.; Lin, Y.S.; Chang, C.H.; Wei, T.Y.; Huang, J.C.; Liao, Z.X.; Lin, C.H.; Chen, C.H. Promising Ta-Ti-Zr-Si metallic glass coating without cytotoxic elements for bio-implant applications. Appl. Surf. Sci. 2018, 427, 485–495. [Google Scholar] [CrossRef]
  176. Chu, J.H.; Lee, J.; Chang, C.C.; Chan, Y.C.; Liou, M.L.; Lee, J.W.; Jang, J.S.C.; Duh, J.G. Antimicrobial characteristics in Cu-containing Zr-based thin film metallic glass. Surf. Coat. Technol. 2014, 259, 87–93. [Google Scholar] [CrossRef]
  177. Peter, W.H.H.; Buchanan, R.A.; Liu, C.T.; Liaw, P.K.; Morrison, M.L.; Horton, J.A.; Carmichael Jr, 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]
  178. Pang, S.; Liu, Y.; Li, H.; Sun, L.; Li, Y.; Zhang, T. New Ti-based Ti–Cu–Zr–Fe–Sn–Si–Ag bulk metallic glass for biomedical applications. J. Alloys Compd. 2014, 625, 323–327. [Google Scholar] [CrossRef]
  179. Tsai, P.H.H.; Lin, Y.Z.; Li, J.B.; Jian, S.R.; Jang, J.S.C.; Li, C.; Chu, J.P.; Huang, J.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]
  180. 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]
  181. Lee, M.; Lee, C.-M.; Lee, K.-R.; Ma, E.; Lee, J.-C. Networked interpenetrating connections of icosahedra: Effects on shear transformations in metallic glass. Acta Mater. 2011, 59, 159–170. [Google Scholar] [CrossRef]
  182. Hui, X.; Fang, H.Z.; Chen, G.L.; Shang, S.L.; Wang, Y.; Qin, J.Y.; Liu, Z.K. Atomic structure of Zr41.2Ti13.8Cu12.5Ni10Be22.5 bulk metallic glass alloy. Acta Mater. 2009, 57, 376–391. [Google Scholar] [CrossRef]
  183. Bernal, J.D. Geometry of the Structure of Monatomic Liquids. Nature 1960, 185, 68–70. [Google Scholar] [CrossRef]
  184. Sheng, H.W.; Luo, W.K.; Alamgir, F.M.; Bai, J.M.; Ma, E. Atomic packing and short-to-medium-range order in metallic glasses. Nature 2006, 439, 419–425. [Google Scholar] [CrossRef] [PubMed]
  185. Gaskell, P.H. A new structural model for amorphous transition metal silicides, borides, phosphides and carbides. J. Non-Cryst. Solids 1979, 32, 207–224. [Google Scholar] [CrossRef]
  186. Miracle, D.B. A structural model for metallic glasses. Nat. Mater. 2004, 3, 697–702. [Google Scholar] [CrossRef]
  187. Zhang, L.; Wu, Y.; Bian, X.; Li, H.; Wang, W.; Wu, S. Short-range and medium-range order in liquid and amorphous Al90Fe5Ce5 alloys. J. Non-Cryst. Solids 2000, 262, 169–176. [Google Scholar] [CrossRef]
  188. Hirotsu, Y.; Matsushita, M.; Ohkubo, T.; Makino, A.; Oikawa, T. Average and local structures of amorphous Pd75Si25 alloy analyzed by modern electron diffraction techniques. Mater. Sci. Eng. A 1997, 226–228, 274–279. [Google Scholar] [CrossRef]
  189. Dubois, J.M.; Le Caer, G. Ordre local et proprietes physiques des verres metalliques riches en fer. Acta Metall. 1984, 32, 2101–2114. [Google Scholar] [CrossRef]
  190. Hufnagel, T.C.; Brennan, S. Short- and medium-range order in (Zr70Cu20Ni1090)xTaxAl10 bulk amorphous alloys. Phys. Rev. B 2003, 67, 014203. [Google Scholar] [CrossRef]
  191. Sietsma, J.; Thijsse, B.J. An investigation of universal medium range order in metallic glasses. J. Non-Cryst. Solids 1991, 135, 146–154. [Google Scholar] [CrossRef]
  192. Wakeda, M.; Shibutani, Y. Icosahedral clustering with medium-range order and local elastic properties of amorphous metals. Acta Mater. 2010, 58, 3963–3969. [Google Scholar] [CrossRef]
  193. Dougherty, G.M.; He, Y.; Shiflet, G.J.; Poon, S.J. Compositional dependence of glass formability in AlNiFeGd amorphous alloys. Scr. Metall. Mater. 1994, 30, 101–106. [Google Scholar] [CrossRef]
  194. Yavari, A.R.; Uriarte, J.-L.; Inoue, A. Volume Effects in Amorphisation by Rapid Solidification and Solid-State Reaction and in Bulk Glass-Forming Alloys. Mater. Sci. Forum 2009, 269–272, 533–540. [Google Scholar] [CrossRef]
  195. Battezzati, L.; Baricco, M. An analysis of volume effects in metallic glass formation. J. Less Common Met. 1988, 145, 31–38. [Google Scholar] [CrossRef]
  196. 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]
  197. Jiang, Q.K.; Wang, X.D.; Nie, X.P.; Zhang, G.Q.; Ma, H.; Fecht, H.-J.; Bendnarcik, J.; Franz, H.; Liu, Y.G.; Cao, Q.P.; et al. Zr–(Cu,Ag)–Al bulk metallic glasses. Acta Mater. 2008, 56, 1785–1796. [Google Scholar] [CrossRef]
  198. Khan, M.M.; Shabib, I.; Haider, W. A combinatorially developed Zr-Ti-Fe-Al metallic glass with outstanding corrosion resistance for implantable medical devices. Scr. Mater. 2019, 162, 223–229. [Google Scholar] [CrossRef]
  199. Liu, J.B.; Zhang, L.; Dong, A.P.; Wang, L.T.; Zeng, Y.W.; Meng, L. Effects of Cr and Zr additions on the microstructure and properties of Cu–6 wt.% Ag alloys. Mater. Sci. Eng. A 2012, 532, 331–338. [Google Scholar] [CrossRef]
  200. Wang, H.-S.; Su, Y.-Z.; Jang, J.S.-C.; Chen, H.-G. A comparison of crystallization behaviors of laser spot welded Zr–Cu–Ag–Al and Zr–Cu–Ni–Al bulk metallic glasses. Mater. Chem. Phys. 2013, 139, 215–219. [Google Scholar] [CrossRef]
  201. Apreutesei, M.; Billard, A.; Steyer, P. Crystallization and hardening of Zr-40 at. % Cu thin film metallic glass: Effects of isothermal annealing. Mater. Des. 2015, 86, 555–563. [Google Scholar] [CrossRef]
  202. Ishii, A.; Iwase, A.; Fukumoto, Y.; Yokoyama, Y.; Konno, T.J.; Hori, F. Effect of thermal annealing on the local structure in ZrCuAl bulk metallic glass. J. Alloys Compd. 2010, 504, S230–S233. [Google Scholar] [CrossRef]
  203. He, G.; Zhang, Z.F.; Löser, W.; Eckert, J.; Schultz, L. Effect of Ta on glass formation, thermal stability and mechanical properties of a Zr52.25Cu28.5Ni4.75Al9.5Ta5 bulk metallic glass. Acta Mater. 2003, 51, 2383–2395. [Google Scholar] [CrossRef]
  204. Inoue, A.; Shibata, T.; Zhang, T. Effect of Additional Elements on Glass Transition Behavior and Glass Formation Tendency of Zr-Al-Cu-Ni Alloys. Mater. Trans. JIM 1995, 36, 1420–1426. [Google Scholar] [CrossRef]
  205. Choi-Yim, H.; Conner, R.D.; Szuecs, F.; Johnson, W.L. Processing, microstructure and properties of ductile metal particulate reinforced Zr57Nb5Al10Cu15.4Ni12.6 bulk metallic glass composites. Acta Mater. 2002, 50, 2737–2745. [Google Scholar] [CrossRef]
  206. Inoue, A.; Zhang, W. Formation, Thermal Stability and Mechanical Properties of Cu-Zr-Al Bulk Glassy Alloys. Mater. Trans. 2005, 43, 2921–2925. [Google Scholar] [CrossRef]
  207. Pang, S.J.; Zhang, T.; Asami, K.; Inoue, A. Synthesis of Fe–Cr–Mo–C–B–P bulk metallic glasses with high corrosion resistance. Acta Mater. 2002, 50, 489–497. [Google Scholar] [CrossRef]
  208. Inoue, A. High Strength Bulk Amorphous Alloys with Low Critical Cooling Rates. Mater. Trans. JIM 1995, 36, 866–875. [Google Scholar] [CrossRef]
  209. Wu, J.; Grosdidier, T.; Allain-Bonasso, N.; Gaffet, E.; Dong, C. On the mechanically induced crystallization of FCC phases by mechanical milling in ZrAlNiCu bulk metallic glasses. J. Alloys Compd. 2010, 504, S264–S266. [Google Scholar] [CrossRef]
  210. Axinte, E. Metallic glasses from ‘alchemy’ to pure science: Present and future of design, processing and applications of glassy metals. Mater. Des. 2012, 35, 518–556. [Google Scholar] [CrossRef]
  211. Telford, M. The case for bulk metallic glass. Mater. Today 2004, 7, 36–43. [Google Scholar] [CrossRef]
  212. Liu, L.; Qiu, C.L.; Huang, C.Y.; Yu, Y.; Huang, H.; Zhang, S.M. Biocompatibility of Ni-free Zr-based bulk metallic glasses. Intermetallics 2009, 17, 235–240. [Google Scholar] [CrossRef]
  213. Liu, C.T.; Heatherly, L.; Horton, J.A.; Easton, D.S.; Carmichael, C.A.; Wright, J.L.; Schneibel, J.H.; Yoo, M.H.; Chen, C.H.; Inoue, A. Test environments and mechanical properties of Zr-base bulk amorphous alloys. Metall. Mater. Trans. A 1998, 29, 1811–1820. [Google Scholar] [CrossRef]
  214. 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]
  215. Liu, Y.; Wang, Y.-M.M.; Pang, H.-F.F.; Zhao, Q.; Liu, L. A Ni-free ZrCuFeAlAg bulk metallic glass with potential for biomedical applications. Acta Biomater. 2013, 9, 7043–7053. [Google Scholar] [CrossRef]
  216. Sun, Y.; Huang, Y.; Fan, H.; Liu, F.; Shen, J.; Sun, J.; Chen, J.J.J. Comparison of mechanical behaviors of several bulk metallic glasses for biomedical application. J. Non-Cryst. Solids 2014, 406, 144–150. [Google Scholar] [CrossRef]
  217. Zhu, S.L.L.; Wang, X.M.M.; Qin, F.X.; Inoue, A.; Qin, F.X.; Inoue, A. A new Ti-based bulk glassy alloy with potential for biomedical application. Mater. Sci. Eng. A 2007, 459, 233–237. [Google Scholar] [CrossRef]
  218. Wang, J.G.; Choi, B.W.; Nieh, T.G.; Liu, C.T. Crystallization and nanoindentation behavior of a bulk Zr–Al–Ti–Cu–Ni amorphous alloy. J. Mater. Res. 2000, 15, 798–807. [Google Scholar] [CrossRef]
  219. Xie, G.; Qin, F.; Zhu, S.; Inoue, A. Ni-free Ti-based bulk metallic glass with potential for biomedical applications produced by spark plasma sintering. Intermetallics 2012, 29, 99–103. [Google Scholar] [CrossRef]
  220. Oak, J.J.; Louzguine-Luzgin, D.V.; Inoue, A. Investigation of glass-forming ability, deformation and corrosion behavior of Ni-free Ti-based BMG alloys designed for application as dental implants. Mater. Sci. Eng. C 2009, 29, 322–327. [Google Scholar] [CrossRef]
  221. Jin, K.; Löffler, J.F. Bulk metallic glass formation in Zr–Cu–Fe–Al alloys. Appl. Phys. Lett. 2005, 86, 241909. [Google Scholar] [CrossRef]
  222. Gu, X.J.; Poon, S.J.; Shiflet, G.J. Mechanical properties of iron-based bulk metallic glasses. J. Mater. Res. 2007, 22, 344–351. [Google Scholar] [CrossRef]
  223. Conner, R.D.; Dandliker, R.B.; Johnson, W.L. Mechanical properties of tungsten and steel fiber reinforced Zr41.25Ti13.75Cu12.5Ni10Be22.5 metallic glass matrix composites. Acta Mater. 1998, 46, 6089–6102. [Google Scholar] [CrossRef]
  224. Qiu, K.Q.; Wang, A.M.; Zhang, H.F.; Ding, B.Z.; Hu, Z.Q. Mechanical properties of tungsten fiber reinforced ZrAlNiCuSi metallic glass matrix composite. Intermetallics 2002, 10, 1283–1288. [Google Scholar] [CrossRef]
  225. Wright, W.; Saha, R.; Nix, W. Deformation Mechanisms of the Zr40Ti14Ni10Cu12Be24 Bulk Metallic Glass. Mater. Trans. 2001, 42, 642–649. [Google Scholar] [CrossRef]
  226. Tang, C.; Li, Y.; Zeng, K. Characterization of mechanical properties of a Zr-based metallic glass by indentation techniques. Mater. Sci. Eng. A 2004, 384, 215–223. [Google Scholar] [CrossRef]
  227. Bei, H.; Lu, Z.P.; George, E.P. Theoretical Strength and the Onset of Plasticity in Bulk Metallic Glasses Investigated by Nanoindentation with a Spherical Indenter. Phys. Rev. Lett. 2004, 93, 125504. [Google Scholar] [CrossRef]
  228. Liu, L.; Chan, K.C. Plastic deformation of Zr-based bulk metallic glasses under nanoindentation. Mater. Lett. 2005, 59, 3090–3094. [Google Scholar] [CrossRef]
  229. Fornell, J.; Pellicer, E.; Steenberge, N.V.; González, S.; Gebert, A.; Suriñach, S.; Baró, M.D.; Sort, J. Improved plasticity and corrosion behavior in Ti–Zr–Cu–Pd metallic glass with minor additions of Nb: An alloy composition intended for biomedical applications. Mater. Sci. Eng. A 2013, 559, 159–164. [Google Scholar] [CrossRef]
  230. Huang, Y.; Chiu, Y.L.; Shen, J.; Chen, J.J.J.; Sun, J. Nanoindentation study of Ti-based metallic glasses. J. Alloys Compd. 2009, 479, 121–128. [Google Scholar] [CrossRef]
  231. Subramanian, B. In vitro corrosion and biocompatibility screening of sputtered Ti40Cu36Pd14Zr10 thin film metallic glasses on steels. Mater. Sci. Eng. C 2015, 47, 48–56. [Google Scholar] [CrossRef]
  232. Kim, Y.C.; Bae, D.H.; Kim, W.T.; Kim, D.H. Glass forming ability and crystallization behavior of Ti-based amorphous alloys with high specific strength. J. Non-Cryst. Solids 2003, 325, 242–250. [Google Scholar] [CrossRef]
  233. Inoue, A.; Zhang, W.; Zhang, T.; Kurosaka, K. High-strength Cu-based bulk glassy alloys in Cu–Zr–Ti and Cu–Hf–Ti ternary systems. Acta Mater. 2001, 49, 2645–2652. [Google Scholar] [CrossRef]
  234. Jiang, W.H.; Liu, F.X.; Wang, Y.D.; Zhang, H.F.; Choo, H.; Liaw, P.K. Comparison of mechanical behavior between bulk and ribbon Cu-based metallic glasses. Mater. Sci. Eng. A 2006, 430, 350–354. [Google Scholar] [CrossRef]
  235. Qin, C.L.; Zhang, W.; Asami, K.; Kimura, H.; Wang, X.M.; Inoue, A. A novel Cu-based BMG composite with high corrosion resistance and excellent mechanical properties. Acta Mater. 2006, 54, 3713–3719. [Google Scholar] [CrossRef]
  236. Park, E.S.; Kim, D.H.; Ohkubo, T.; Hono, K. Enhancement of glass forming ability and plasticity by addition of Nb in Cu–Ti–Zr–Ni–Si bulk metallic glasses. J. Non-Cryst. Solids 2005, 351, 1232–1238. [Google Scholar] [CrossRef]
  237. Cheung, T.L.; Shek, C.H. Thermal and mechanical properties of Cu–Zr–Al bulk metallic glasses. J. Alloys Compd. 2007, 434–435, 71–74. [Google Scholar] [CrossRef]
  238. Jia, P.; Guo, H.; Li, Y.; Xu, J.; Ma, E. A new Cu–Hf–Al ternary bulk metallic glass with high glass forming ability and ductility. Scr. Mater. 2006, 54, 2165–2168. [Google Scholar] [CrossRef]
  239. Bian, Z.; Kato, H.; Qin, C.; Zhang, W.; Inoue, A. Cu–Hf–Ti–Ag–Ta bulk metallic glass composites and their properties. Acta Mater. 2005, 53, 2037–2048. [Google Scholar] [CrossRef]
  240. Chen, Q.J.; Shen, J.; Zhang, D.L.; Fan, H.B.; Sun, J.F. Mechanical performance and fracture behavior of Fe41Co7Cr15Mo14Y2C15B6 bulk metallic glass. J. Mater. Res. 2007, 22, 358–363. [Google Scholar] [CrossRef]
  241. Fornell, J.; González, S.; Rossinyol, E.; Suriñach, S.; Baró, M.D.; Louzguine-Luzgin, D.V.; Perepezko, J.H.; Sort, J.; Inoue, A. Enhanced mechanical properties due to structural changes induced by devitrification in Fe–Co–B–Si–Nb bulk metallic glass. Acta Mater. 2010, 58, 6256–6266. [Google Scholar] [CrossRef]
  242. Shen, B.; Chang, C.; Inoue, A. Formation, ductile deformation behavior and soft-magnetic properties of (Fe,Co,Ni)–B–Si–Nb bulk glassy alloys. Intermetallics 2007, 15, 9–16. [Google Scholar] [CrossRef]
  243. Inoue, A.; Shen, B.L.; Chang, C.T. Super-high strength of over 4000 MPa for Fe-based bulk glassy alloys in [(Fe1−xCox)0.75B0.2Si0.05]96Nb4 system. Acta Mater. 2004, 52, 4093–4099. [Google Scholar] [CrossRef]
  244. Gu, X.J.; Poon, S.J.; Shiflet, G.J.; Widom, M. Ductility improvement of amorphous steels: Roles of shear modulus and electronic structure. Acta Mater. 2008, 56, 88–94. [Google Scholar] [CrossRef]
  245. Xu, F.; Ding, Y.H.; Deng, X.H.; Zhang, P.; Long, Z.L. Indentation size effects in the nano- and micro-hardness of a Fe-based bulk metallic glass. Phys. B Condens. Matter 2014, 450, 84–89. [Google Scholar] [CrossRef]
  246. Schuster, B.E.; Wei, Q.; Hufnagel, T.C.; Ramesh, K.T. Size-independent strength and deformation mode in compression of a Pd-based metallic glass. Acta Mater. 2008, 56, 5091–5100. [Google Scholar] [CrossRef]
  247. Choi-Yim, H.; Xu, D.; Johnson, W.L. Ni-based bulk metallic glass formation in the Ni–Nb–Sn and Ni–Nb–Sn–X (X = B, Fe, Cu) alloy systems. Appl. Phys. Lett. 2003, 82, 1030–1032. [Google Scholar] [CrossRef]
  248. 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]
  249. Li, H.F.; Xie, X.H.; Zhao, K.; Wang, Y.B.; Zheng, Y.F.; Wang, W.H.; Qin, L. In vitro and in vivo studies on biodegradable CaMgZnSrYb high-entropy bulk metallic glass. Acta Biomater. 2013, 9, 8561–8573. [Google Scholar] [CrossRef]
  250. Inoue, A.; Ohtera, K.; Tsai, A.-P.; Masumoto, T. Aluminum-Based Amorphous Alloys with Tensile Strength above 980 MPa (100 kg/mm2). Jpn. J. Appl. Phys. 1988, 27, L479–L482. [Google Scholar] [CrossRef]
  251. Gilbert, C.J.; Ritchie, R.O.; Johnson, W.L. Fracture toughness and fatigue-crack propagation in a Zr–Ti–Ni–Cu–Be bulk metallic glass. Appl. Phys. Lett. 1997, 71, 476–478. [Google Scholar] [CrossRef]
  252. Kawamura, Y.; Kato, H.; Inoue, A.; Masumoto, T. Full strength compacts by extrusion of glassy metal powder at the supercooled liquid state. Appl. Phys. Lett. 1995, 67, 2008–2010. [Google Scholar] [CrossRef]
  253. Choi-Yim, H.; Busch, R.; Köster, U.; Johnson, W.L. Synthesis and characterization of particulate reinforced Zr57Nb5Al10Cu15.4Ni12.6 bulk metallic glass composites. Acta Mater. 1999, 47, 2455–2462. [Google Scholar] [CrossRef]
  254. Fan, C.; Li, C.; Inoue, A.; Haas, V. Deformation behavior of Zr-based bulk nanocrystalline amorphous alloys. Phys. Rev. B 2000, 61, R3761–R3763. [Google Scholar] [CrossRef]
  255. Leonhard, A.; Xing, L.Q.; Heilmaier, M.; Gebert, A.; Eckert, J.; Schultz, L. Effect of crystalline precipitations on the mechanical behavior of bulk glass forming Zr-based alloys. Nanostruct. Mater. 1998, 10, 805–817. [Google Scholar] [CrossRef]
  256. Fan, C.; Takeuchi, A.; Inoue, A. Preparation and Mechanical Properties of Zr-based Bulk Nanocrystalline Alloys Containing Compound and Amorphous Phases. Mater. Trans. JIM 1999, 40, 42–51. [Google Scholar] [CrossRef]
  257. Eckert, J.; Mattern, N.; Zinkevitch, M.; Seidel, M. Crystallization behavior and phase formation in Zr–Al–Cu–Ni metallic glass containing oxygen. Mater. Trans. JIM 1998, 39, 623–632. [Google Scholar] [CrossRef]
  258. Liu, Z.Q.; Liu, G.; Qu, R.T.; Zhang, Z.F.; Wu, S.J.; Zhang, T. Microstructural percolation assisted breakthrough of trade-off between strength and ductility in CuZr-based metallic glass composites. Sci. Rep. 2015, 4, 4167. [Google Scholar] [CrossRef]
  259. Eckert, J.; Das, J.; Kim, K.B.; Baier, F.; Tang, M.B.; Wang, W.H.; Zhang, Z.F. High strength ductile Cu-base metallic glass. Intermetallics 2006, 14, 876–881. [Google Scholar] [CrossRef]
  260. Pauly, S.; Liu, G.; Wang, G.; Kühn, U.; Mattern, N.; Eckert, J.H. Microstructural heterogeneities governing the deformation of Cu47.5Zr47.5Al5 bulk metallic glass composites. Acta Mater. 2009, 57, 5445–5453. Available online: https://www.sciencedirect.com/science/article/pii/S1359645409004765 (accessed on 21 June 2023). [CrossRef]
  261. Wu, Y.; Xiao, Y.; Chen, G.; Liu, C.T.; Lu, Z. Bulk Metallic Glass Composites with Transformation-Mediated Work-Hardening and Ductility. Adv. Mater. 2010, 22, 2770–2773. [Google Scholar] [CrossRef]
  262. Wu, Y.; Song, W.L.; Zhang, Z.Y.; Hui, X.; Ma, D.; Wang, X.; Shang, X.; Lu, Z. Relationship between composite structures and compressive properties in CuZr-based bulk metallic glass system. Chin. Sci. Bull. 2011, 56, 3960–3964. [Google Scholar] [CrossRef]
  263. Liu, Z.; Li, R.; Liu, G.; Su, W.; Wang, H.; Li, Y.; Shi, M.; Luo, X.; Wu, G.; Zhang, T. Microstructural tailoring and improvement of mechanical properties in CuZr-based bulk metallic glass composites. Acta Mater. 2012, 60, 6–7. [Google Scholar] [CrossRef]
  264. Aditya, A.V.; Arora, H.S.; Mukherjee, S. Corrosion behavior of ZrTiCuNiBe bulk metallic glass subjected to friction stir processing. J. Non-Cryst. Solids 2015, 425, 124–129. [Google Scholar] [CrossRef]
  265. Jabed, A.; Khan, M.M.; Camiller, J.; Greenlee-Wacker, M.; Haider, W.; Shabib, I. Property optimization of Zr-Ti-X (X = Ag, Al) metallic glass via combinatorial development aimed at prospective biomedical application. Surf. Coat. Technol. 2019, 372, 278–287. [Google Scholar] [CrossRef]
  266. Morrison, M.L.L.; Buchanan, R.A.A.; Peker, A.; Peter, W.H.H.; Horton, J.A.A.; Liaw, P.K.K. Cyclic-anodic-polarization studies of a Zr41.2Ti 13.8Ni10Cu12.5Be22.5 bulk metallic glass. Intermetallic 2004, 12, 1177–1181. [Google Scholar] [CrossRef]
  267. Jabed, A.; Kumar, G. Effect of cryogenic cycling and above-Tg annealing on the corrosion behavior of Zr-Cu-Ti-Be metallic glass. J. Non-Cryst. Solids 2023, 608, 122260. [Google Scholar] [CrossRef]
  268. Al-Mamun, N.S.; Jabed, A.; Noor-A-Alam, M.; Haider, W.; Shabib, I. Electrochemical response in physiological environments and mechanical properties of Ti51Si24Ni14Nb11 and Fe40Ti30Al18Si12 thin film metallic glasses. J. Non-Cryst. Solids 2022, 575, 121209. [Google Scholar] [CrossRef]
  269. Masumoto, T.; Hashimoto, K. Chemical properties of amorphous metals. Annu. Rev. Mater. Sci. 1978, 8, 215–233. [Google Scholar] [CrossRef]
  270. Kou, H.; Li, Y.; Zhang, T.; Li, J.; Li, J. Electrochemical corrosion properties of Zr- and Ti-based bulk metallic glasses. Trans. Nonferrous Met. Soc. China 2011, 21, 552–557. [Google Scholar] [CrossRef]
  271. Jiang, W.H.; Jiang, F.; Green, B.A.; Liu, F.X.; Liaw, P.K.; Choo, H.; Qiu, Q. Electrochemical corrosion behavior of a Zr-based bulk-metallic glass. Appl. Phys. Lett. 2007, 91, 041904. [Google Scholar] [CrossRef]
  272. Morrison, M.L.L.; Buchanan, R.A.A.; Peker, A.; Liaw, P.K.K.; Horton, J.A.A. Electrochemical behavior of a Ti-based bulk metallic glass. J. Non-Cryst. Solids 2007, 353, 2115–2124. [Google Scholar] [CrossRef]
  273. Jabed, A.; Rahman, Z.U.; Khan, M.M.; Haider, W.; Shabib, I. Combinatorial Development and In Vitro Characterization of the Quaternary Zr–Ti–X–Y (X–Y = Cu–Ag/Co–Ni) Metallic Glass for Prospective Bioimplants. Adv. Eng. Mater. 2019, 21, 1900726. [Google Scholar] [CrossRef]
  274. Stansbury, E.E.; Buchanan, R.A. Fundamentals of Electrochemical Corrosion—Ele Eugene Stansbury, Robert Angus Buchanan; ASM International: Materials Park, OH, USA, 2000. [Google Scholar]
  275. 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]
  276. Luborsky, F.E. Amorphous metallic alloys. In Amorphous Metallic Alloys; Butterworths: London, UK, 1983. [Google Scholar]
  277. Wang, Y.B.; Li, H.F.; Cheng, Y.; Wei, S.C.; Zheng, Y.F. Corrosion performances of a Nickel-free Fe-based bulk metallic glass in simulated body fluids. Electrochem. Commun. 2009, 11, 2187–2190. [Google Scholar] [CrossRef]
  278. Huang, L.; Wang, G.; Qiao, D.; Liaw, P.K.; Pang, S.; Wang, J. Corrosion-fatigue study of a Zr-based bulk-metallic glass in a physiologically relevant environment. J. Alloys Compd. 2010, 504, S159–S162. [Google Scholar] [CrossRef]
  279. Lee, J.; Liou, M.L.; Duh, J.G. The development of a Zr-Cu-Al-Ag-N thin film metallic glass coating in pursuit of improved mechanical, corrosion, and antimicrobial property for bio-medical application. Surf. Coat. Technol. 2017, 310, 214–222. [Google Scholar] [CrossRef]
  280. Fornell, J.; Pellicer, E.; Steenberge, N.V.; 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]
  281. Huang, H.-H.; Sun, Y.-S.; Wu, C.-P.; Liu, C.-F.; Liaw, P.K.; Kai, W. Corrosion resistance and biocompatibility of Ni-free Zr-based bulk metallic glass for biomedical applications. Intermetallics 2012, 30, 139–143. [Google Scholar] [CrossRef]
  282. Jayaraj, J.; Park, J.M.; Gostin, P.F.; Fleury, E.; Gebert, A.; Schultz, L. Nano-porous surface states of Ti–Y–Al–Co phase separated metallic glass. Intermetallics 2009, 17, 1120–1123. [Google Scholar] [CrossRef]
  283. Chang, J.-C.; Lee, J.-W.; Lou, B.-S.; Li, C.-L.; Chu, J.P. Effects of tungsten contents on the microstructure, mechanical and anticorrosion properties of Zr–W–Ti thin film metallic glasses. Thin Solid Film. 2015, 584, 253–256. [Google Scholar] [CrossRef]
  284. Hua, N.; Huang, L.; Wang, J.; Cao, Y.; He, W.; Pang, S.; Zhang, T. Corrosion behavior and in vitro biocompatibility of Zr–Al–Co–Ag bulk metallic glasses: An experimental case study. J. Non-Cryst. Solids 2012, 358, 1599–1604. [Google Scholar] [CrossRef]
  285. Qiu, C.L.; Chen, Q.; Liu, L.; Chan, K.C.; Zhou, J.X.; Chen, P.P.; Zhang, S.M. A novel Ni-free Zr-based bulk metallic glass with enhanced plasticity and good biocompatibility. Scr. Mater. 2006, 55, 605–608. [Google Scholar] [CrossRef]
  286. Hiromoto, S.; Tsai, A.-P.; Sumita, M.; Hanawa, T. Effect of chloride ion on the anodic polarization behavior of the Zr65Al7.5Ni10Cu17.5 amorphous alloy in phosphate buffered solution. Corros. Sci. 2000, 42, 1651–1660. [Google Scholar] [CrossRef]
  287. Oak, J.-J.; Louzguine-Luzgin, D.V.; Inoue, A. Fabrication of Ni-free Ti-based bulk-metallic glassy alloy having potential for application as biomaterial, and investigation of its mechanical properties, corrosion, and crystallization behavior. J. Mater. Res. 2007, 22, 1346–1353. [Google Scholar] [CrossRef]
  288. Wang, T.; Wu, Y.D.; Si, J.J.; Cai, Y.H.; Chen, X.H.; Hui, X.D. Novel Ti-based bulk metallic glasses with superior plastic yielding strength and corrosion resistance. Mater. Sci. Eng. A 2015, 642, 297–303. [Google Scholar] [CrossRef]
  289. Nnamchi, P.S.; Obayi, C.S.; Todd, I.; Rainforth, M.W. Mechanical and electrochemical characterization of new Ti–Mo–Nb–Zr alloys for biomedical applications. J. Mech. Behav. Biomed. Mater. 2016, 60, 68–77. [Google Scholar] [CrossRef]
  290. 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]
  291. Mukherjee, S.; Schroers, J.; Johnson, W.L.; Rhim, W.-K.K. Influence of Kinetic and Thermodynamic Factors on the Glass-Forming Ability of Zirconium-Based Bulk Amorphous Alloys. Phys. Rev. Lett. 2005, 94, 245501. [Google Scholar] [CrossRef]
  292. Shen, J.; Chen, Q.; Sun, J.; Fan, H.; Wang, G. Exceptionally high glass-forming ability of an FeCoCrMoCBY alloy. Appl. Phys. Lett. 2005, 86, 151907. [Google Scholar] [CrossRef]
  293. Wang, P. Corrosion Behaviour of Zirconium Alloys in High Temperature Aqueous Environment by Electrochemical Impedance Spectroscopy October; The University of Manchester Library: Manchester, UK, 2011; pp. 1–217. Available online: https://www.escholar.manchester.ac.uk/uk-ac-man-scw:138826 (accessed on 21 June 2023).
  294. Mareci, D.; Chelariu, R.; Gordin, D.-M.; Ungureanu, G.; Gloriant, T. Comparative corrosion study of Ti–Ta alloys for dental applications. Acta Biomater. 2009, 5, 3625–3639. [Google Scholar] [CrossRef]
  295. Wang, Y.; Jiang, S.L.; Zheng, Y.G.; Ke, W.; Sun, W.H.; Wang, J.Q. Electrochemical behaviour of Fe-based metallic glasses in acidic and neutral solutions. Corros. Sci. 2012, 63, 159–173. [Google Scholar] [CrossRef]
  296. Liu, L.; Qiu, C.L.L.; Chen, Q.; Zhang, S.M.M. Corrosion behavior of Zr-based bulk metallic glasses in different artificial body fluids. J. Alloys Compd. 2006, 425, 268–273. [Google Scholar] [CrossRef]
  297. 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] [PubMed]
  298. Wada, T.; Qin, F.; Wang, X.; Yoshimura, M.; Inoue, A.; Sugiyama, N.; Ito, R.; Matsushita, N. Formation and bioactivation of Zr-Al-Co bulk metallic glasses. J. Mater. Res. 2009, 24, 2941–2948. [Google Scholar] [CrossRef]
  299. Liu, B.; Liu, L. The effect of microalloying on thermal stability and corrosion resistance of Cu-based bulk metallic glasses. Mater. Sci. Eng. A 2006, 415, 286–290. [Google Scholar] [CrossRef]
  300. Apreutesei, M.; Boissy, C.; Mary, N.; Yazdi, M.A.P.; Billard, A.; Steyer, P. Binary Zr–Ni/Co metallic glass films: Role of the structural state on their durability. Acta Mater. 2015, 89, 305–314. [Google Scholar] [CrossRef]
  301. Yang, Y.J.; Jin, Z.S.; Ma, X.Z.; Zhang, Z.P.; Zhang, H.; Ma, M.Z.; Zhang, X.Y.; Li, G.; Liu, R.P. Comparison of corrosion behaviors between Ti-based bulk metallic glasses and its composites. J. Alloys Compd. 2018, 750, 757–764. [Google Scholar] [CrossRef]
  302. Liu, L.; Qiu, C.L.L.; Chen, Q.; Chan, K.C.C.; Zhang, S.M.M. Deformation behavior, corrosion resistance, and cytotoxicity of Ni-free Zr-based bulk metallic glasses. J. Biomed. Mater. Res. Part A 2008, 86A, 160–169. [Google Scholar] [CrossRef]
  303. Zhang, J.; Shi, P.; Chang, A.; Zhao, T.; Li, W.; Chang, C.; Jia, J.; Wong, Q.; You, F.; Feng, D.; et al. Glass-forming ability, thermal stability, mechanical and electrochemical behavior of Al-Ce-TM (TM = Ti, Cr, Mn, Fe, Co, Ni and Cu) amorphous alloys. J. Non-Cryst. Solids X 2019, 1, 100005. [Google Scholar] [CrossRef]
  304. Diyatmika, W.; Chu, J.P.; Kacha, B.T.; Yu, C.-C.C.; Lee, C.-M.M. Thin film metallic glasses in optoelectronic, magnetic, and electronic applications: A recent update. Curr. Opin. Solid State Mater. Sci. 2015, 19, 95–106. [Google Scholar] [CrossRef]
  305. Nishibe, Y.; Yamadera, H.; Ohta, N.; Tsukada, K.; Nonomura, Y. Thin film magnetic field sensor utilizing Magneto Impedance effect. Sens. Actuators A Phys. 2000, 82, 155–160. [Google Scholar] [CrossRef]
  306. Hristoforou, E.; Hristoforou, E. Magnetic effects in physical sensor design magnetic effects in physical sensor design and development. J. Optoelectron. Adv. Mater. 2002, 4, 245–260. [Google Scholar]
  307. Pookat, G.; Thomas, H.; Thomas, S.; Al-Harthi, S.H.; Raghavan, L.; Al-Omari, I.A.; Sakthikumar, D.; Ramanujan, R.V.; Anantharagman, M.R. Evolution of structural and magnetic properties of Co–Fe based metallic glass thin films with thermal annealing. Surf. Coat. Technol. 2013, 236, 246–251. [Google Scholar] [CrossRef]
  308. Crocker, A.G.; Elvidge, A.M.; Flewitt, P.E.J. Structure and properties of polycrystalline materials. Phys. Scr. 1987, 1987, 344–349. [Google Scholar] [CrossRef]
  309. Jin, Y.M.; Wang, Y.U.; Kazaryan, A.; Wang, Y.; Laughlin, D.E.; Khachaturyan, A.G. Magnetic structure and hysteresis in hard magnetic nanocrystalline film: Computer simulation. J. Appl. Phys. 2002, 92, 6172–6181. [Google Scholar] [CrossRef]
  310. Chu, J.P.; Lo, C.-T.; Fang, Y.-K.; Han, B.-S. On annealing-induced amorphization and anisotropy in a ferromagnetic Fe-based film: A magnetic and property study. Appl. Phys. Lett. 2006, 88, 012510. [Google Scholar] [CrossRef]
  311. Jiao, Z.B.; Li, H.X.; Gao, J.E.; Wu, Y.; Lu, Z.P. Effects of alloying elements on glass formation, mechanical and soft-magnetic properties of Fe-based metallic glasses. Intermetallics 2011, 19, 1502–1508. [Google Scholar] [CrossRef]
  312. Mehrizi, S.; Sohi, M.H.; Ebrahimi, S.A.S. Study of microstructure and magnetic properties of electrodeposited nanocrystalline CoFeNiCu thin films. Surf. Coat. Technol. 2011, 205, 4757–4763. [Google Scholar] [CrossRef]
  313. Zhu, M.; Zhang, M.; Yao, L.; Nan, R.; Jian, Z.; Chang, F. Effect of Mo substitution for Nb on the glass-forming ability, magnetic properties, and electrical resistivity in Fe80(Nb1−xMox)5B15 (x = 0–0.75) amorphous ribbons. Vacuum 2019, 163, 368–372. [Google Scholar] [CrossRef]
  314. Shen, B.; Inoue, A. Bulk Glassy Fe-Ga-P-C-B-Si Alloys with High Glass-Forming Ability, High Saturation Magnetization and Good Soft Magnetic Properties The effect of Si addition on the glass-forming ability and magnetic properties for the Fe 77 Ga. Mater. Trans. 2002, 43, 1235–1239. [Google Scholar] [CrossRef]
  315. Lin, C.-Y.; Tien, H.-Y.; Chin, T.-S. Soft magnetic ternary iron-boron-based bulk metallic glasses. Appl. Phys. Lett. 2005, 86, 162501. [Google Scholar] [CrossRef]
  316. Inoue, A.; Shen, B.L.; Koshiba, H.; Kato, H.; Yavari, A.R. Ultra-high strength above 5000 MPa and soft magnetic properties of Co–Fe–Ta–B bulk glassy alloys. Acta Mater. 2004, 52, 1631–1637. [Google Scholar] [CrossRef]
  317. Inoue, A. Bulk amorphous and nanocrystalline alloys with high functional properties. Mater. Sci. Eng. A 2001, 304–306, 1–10. [Google Scholar] [CrossRef]
  318. Mizushima, T.; Ikarashi, K.; Makino, A.; Inoue, A. Structure and magnetic properties of bulk Fe-Al-Ga-P-C-B-Si glassy alloys in a ringed form prepared by copper mold casting. IEEE Trans. Magn. 1999, 35, 3361–3363. [Google Scholar] [CrossRef]
  319. Piccin, R.; Tiberto, P.; Chiriac, H.; Baricco, M. Magnetic properties and power losses in Fe–Co-based bulk metallic glasses. J. Magn. Magn. Mater. 2008, 320, e806–e809. [Google Scholar] [CrossRef]
  320. Inoue, A.; Gook, J.S. Fe-Based Ferromagnetic Glassy Alloys with Wide Supercooled Liquid Region. Mater. Trans. JIM 1995, 36, 1180–1183. [Google Scholar] [CrossRef]
  321. Shen, B.; Inoue, A.; Chang, C. Superhigh strength and good soft-magnetic properties of (Fe,Co)–B–Si–Nb bulk glassy alloys with high glass-forming ability. Appl. Phys. Lett. 2004, 85, 4911–4913. [Google Scholar] [CrossRef]
  322. Makino, A.; Kubota, T.; Makabe, M.; Chang, C.T.; Inoue, A. FeSiBP metallic glasses with high glass-forming ability and excellent magnetic properties. Mater. Sci. Eng. B 2008, 148, 166–170. [Google Scholar] [CrossRef]
  323. Gao, J.E.; Li, H.X.; Jiao, Z.B.; Wu, Y.; Chen, Y.H.; Yu, T.; Lu, P. Effects of nanocrystal formation on the soft magnetic properties of Fe-based bulk metallic glasses. Appl. Phys. Lett. 2011, 99, 052504. [Google Scholar] [CrossRef]
  324. Zhang, M.; Kong, F.; Wang, A.; Chang, C.; Shen, B. Soft magnetic properties of bulk FeCoMoPCBSi glassy core prepared by copper mold casting. J. Appl. Phys. 2012, 111, 07A312. [Google Scholar] [CrossRef]
  325. Kong, F.; Wang, A.; Fan, X.; Men, H.; Shen, B.; Xie, G.; Makino, A.; Inoue, A. High Bs Fe84-xSi4B8P 4Cux (x = 0–1.5) nanocrystalline alloys with excellent magnetic softness. J. Appl. Phys. 2011, 109, 07A303. [Google Scholar] [CrossRef]
  326. Yoshizawa, Y.; Oguma, S.; Yamauchi, K. New Fe--based soft magnetic alloys composed of ultrafine grain structure. J. Appl. Phys. 1988, 64, 6044–6046. [Google Scholar] [CrossRef]
  327. Hasegawa, R. Advances in amorphous and nanocrystalline magnetic materials. J. Magn. Magn. Mater. 2006, 304, 187–191. [Google Scholar] [CrossRef]
  328. Quintana, P.; Amano, E.; Valenzuela, R.; Irvine, J.T.S. Effects of nanocrystallization upon the soft magnetic properties of Co--based amorphous alloys. J. Appl. Phys. 1994, 75, 6940–6942. [Google Scholar] [CrossRef]
  329. Tsai, C.S.; Yang, W.J.; Leu, M.S.; Lin, C.S. High frequency characteristics of annealed Co-base amorphous alloy ribbons. J. Appl. Phys. 1991, 70, 5846–5848. [Google Scholar] [CrossRef]
  330. Wang, Q.; Zhu, C.L.; Dong, C.; Qiang, J.B.; Zhang, W.; Inoue, A. Formation and soft magnetic properties of Co (-Fe)-Si-B-Nb bulk metallic glasses in relation to clusters. J. Phys. Conf. Ser. 2008, 98, 012017. [Google Scholar] [CrossRef]
  331. Sun, W.S.; Liang, X.B.; Kulik, T. Formation and magnetic properties of Co–Fe-based bulk metallic glasses with supercooled liquid region. J. Magn. Magn. Mater. 2006, 299, 492–495. [Google Scholar] [CrossRef]
  332. Sun, W.S.; Kulik, T.; Liang, X.B.; Ferenc, J. Thermal stability and magnetic properties of Co-Fe-Hf-Ti-Mo-B bulk metallic glass. Intermetallics 2006, 14, 1066–1068. [Google Scholar] [CrossRef]
  333. Lachowicz, H.K.; Kulik, T.; Żuberek, R.; Małkiński, L.; Kuźmiński, M.; Ślawska-Waniewska, A.; Muñoz, J.S. Tailoring soft and hard magnets by annealing Co-based metallic glass. J. Magn. Magn. Mater. 1998, 190, 267–276. [Google Scholar] [CrossRef]
  334. Chiriac, H.; Óvári, T.A. Amorphous glass-covered magnetic wires: Preparation, properties, applications. Prog. Mater. Sci. 1996, 40, 333–407. [Google Scholar] [CrossRef]
  335. Inoue, A.; Shen, B.; Koshiba, H.; Kato, H.; Yavari, A.R. Cobalt-based bulk glassy alloy with ultrahigh strength and soft magnetic properties. Nat. Mater. 2003, 2, 661–663. [Google Scholar] [CrossRef] [PubMed]
  336. Chu, J.P.; Jang, J.S.C.; Huang, J.C.; Chou, H.S.; Yang, Y.; Ye, J.C.; Wang, Y.C.; Lee, J.W.; Liu, F.X.; Liaw, P.K.; et al. Thin film metallic glasses: Unique properties and potential applications. Thin Solid Films 2012, 520, 5097–5122. [Google Scholar] [CrossRef]
  337. Chang, C.M.; Yang, C.J.; Wang, K.-K.; Liu, J.-K.; Hsu, J.H.; Huang, J.C. On the reflectivity and antibacterial/antifungal responses of Al-Ni-Y optical thin film metallic glass composites. Surf. Coat. Technol. 2017, 327, 75–82. [Google Scholar] [CrossRef]
  338. Lin, B.; Mu, R.; Yang, L.; Bian, X. Antibacterial effect of metallic glasses. Chin. Sci. Bull. 2012, 57, 1069–1072. [Google Scholar] [CrossRef]
  339. Donlan, R.M. Biofilms: Microbial life on surfaces. Emerg. Infect. Dis. 2002, 8, 881–890. [Google Scholar] [CrossRef]
  340. Lemire, J.A.; Harrison, J.J.; Turner, R.J. Antimicrobial activity of metals: Mechanisms, molecular targets and applications. Nat. Rev. Microbiol. 2013, 11, 371–384. [Google Scholar] [CrossRef]
  341. Chiang, P.T.; Chen, G.J.; Jian, S.R.; Shih, Y.H.; Jang, J.S.C.; Lai, C.H. Surface antimicrobial effects of Zr61Al7.5Ni10Cu17.5Si4thin film metallic glasses on Escherichia coli, Staphylococcus aureus, Pseudomonas aeruginosa, Acinetobacter baumannii and Candida albicans. Fooyin J. Health Sci. 2010, 2, 12–20. [Google Scholar] [CrossRef]
  342. Niinomi, M.; Nakai, M.; Hieda, J. Development of new metallic alloys for biomedical applications. Acta Biomater. 2012, 8, 3888–3903. [Google Scholar] [CrossRef]
  343. Huiskes, R.I.K.; Weinans, H.; Van Rietbergen, B. The Relationship Between Stress Shielding and Bone Resorption Around Total Hip Stems and the Effects of Flexible Materials. Clin. Orthop. Relat. Res. 1992, 274, 124–134. [Google Scholar] [CrossRef]
  344. Wang, Y.B.; Zheng, Y.F.; Wei, S.C.; Li, M. In vitro study on Zr-based bulk metallic glasses as potential biomaterials. J. Biomed. Mater. Res. Part B Appl. Biomater. 2011, 96B, 34–46. [Google Scholar] [CrossRef]
  345. Khan, M.M.; Deen, K.M.; Haider, W. Combinatorial development and assessment of a Zr-based metallic glass for prospective biomedical applications. J. Non-Cryst. Solids 2019, 523, 119544. [Google Scholar] [CrossRef]
  346. Wang, Y.B.; Li, H.F.; Zheng, Y.F.; Li, M. Corrosion performances in simulated body fluids and cytotoxicity evaluation of Fe-based bulk metallic glasses. Mater. Sci. Eng. C 2012, 32, 599–606. [Google Scholar] [CrossRef]
  347. Wang, Y.B.; Li, H.F.; Cheng, Y.; Zheng, Y.F.; 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]
  348. Oak, J.-J.; Hwang, G.-W.; Park, Y.-H.; Kimura, H.; Yoon, S.-Y.; Inoue, A. Characterization of Surface Properties, Osteoblast Cell Culture in Vitro and Processing with Flow-Viscosity of Ni-Free Ti-Based Bulk Metallic Glass for Biomaterials. J. Biomech. Sci. Eng. 2009, 4, 384–391. [Google Scholar] [CrossRef]
  349. Xie, K.-F.; Yao, K.-F.; Huang, T.-Y. A Ti-based bulk glassy alloy with high strength and good glass forming ability. Intermetallics 2010, 18, 1837–1841. [Google Scholar] [CrossRef]
  350. Huang, L.; Yokoyama, Y.; Wu, W.; Liaw, P.K.; Pang, S.; Inoue, A.; Zhang, T.; He, W. Ni-free Zr-Cu-Al-Nb-Pd bulk metallic glasses with different Zr/Cu ratios for biomedical applications. J. Biomed. Mater. Res. Part B Appl. Biomater. 2012, 100B, 1472–1482. [Google Scholar] [CrossRef]
  351. He, W.; Chuang, A.; Cao, Z.; Liaw, P.K. Biocompatibility Study of Zirconium-Based Bulk Metallic Glasses for Orthopedic Applications. Metall. Mater. Trans. A 2010, 41, 1726–1734. [Google Scholar] [CrossRef]
  352. Blanquer, A.; Pellicer, E.; Hynowska, A.; Barrios, L.; Ibáñez, E.; Baró, M.D.; Sort, J.; Nogués, C. In vitro biocompatibility assessment of Ti40Cu38Zr10Pd12 bulk metallic glass. J. Mater. Sci. Mater. Med. 2014, 25, 163–172. [Google Scholar] [CrossRef]
  353. Huang, X.B.; Lin, N.M.; Hang, R.Q.; Tang, B.; Qiao, J.W. Bio-Properties of Zr-Based BMGMC as Potential Hard Tissue Implants. Mater. Sci. Forum 2013, 745–746, 754–760. [Google Scholar] [CrossRef]
  354. Sun, Y.; Huang, Y.; Fan, H.; Wang, Y.; Ning, Z.; Liu, F.; Feng, D.; Jin, X.; Shen, J.; Sun, J.; et al. In vitro and in vivo biocompatibility of an Ag-bearing Zr-based bulk metallic glass for potential medical use. J. Non-Cryst. Solids 2015, 419, 82–91. [Google Scholar] [CrossRef]
  355. Li, H.F.; Zheng, Y.F.; Xu, F.; Jiang, J.Z. In vitro investigation of novel Ni free Zr-based bulk metallic glasses as potential biomaterials. Mater. Lett. 2012, 75, 74–76. [Google Scholar] [CrossRef]
  356. Li, J.; Shi, L.-L.; Zhu, Z.; He, Q.; Ai, H.; Xu, J. Zr61Ti2Cu25Al12 metallic glass for potential use in dental implants: Biocompatibility assessment by in vitro cellular responses. Mater. Sci. Eng. C 2013, 33, 2113–2121. [Google Scholar] [CrossRef] [PubMed]
  357. Ramya, M.; Sarwat, S.G.; Udhayabanu, V.; Subramanian, S.; Raj, B.; Ravi, K.R. Role of partially amorphous structure and alloying elements on the corrosion behavior of Mg–Zn–Ca bulk metallic glass for biomedical applications. Mater. Des. 2015, 86, 829–835. [Google Scholar] [CrossRef]
  358. Gu, X.; Shiflet, G.J.; Guo, F.Q.; Poon, S.J. Mg–Ca–Zn Bulk Metallic Glasses with High Strength and Significant Ductility. J. Mater. Res. 2005, 20, 1935–1938. [Google Scholar] [CrossRef]
  359. Zberg, B.; Uggowitzer, P.J.; Löffler, J.F. MgZnCa glasses without clinically observable hydrogen evolution for biodegradable implants. Nat. Mater. 2009, 8, 887–891. [Google Scholar] [CrossRef]
  360. Gu, X.; Zheng, Y.; Zhong, S.; Xi, T.; Wang, J.; Wang, W. Corrosion of, and cellular responses to Mg-Zn-Ca bulk metallic glasses. Biomaterials 2009, 31, 1093–1103. [Google Scholar] [CrossRef]
  361. Khan, M.M.; Rahman, Z.U.; Deen, K.M.; Shabib, I.; Haider, W. Sputtered Mg100-xZnx (0 ≤ x ≤ 100) systems as anode materials for a biodegradable battery aimed for transient bioelectronics. Electrochim. Acta 2020, 329, 135129. [Google Scholar] [CrossRef]
  362. Cao, J.D.; Kirkland, N.T.; Laws, K.J.; Birbilis, N.; Ferry, M. Ca–Mg–Zn bulk metallic glasses as bioresorbable metals. Acta Biomater. 2012, 8, 2375–2383. [Google Scholar] [CrossRef]
  363. Senkov, O.N.; Miracle, D.B.; Keppens, V.; Liaw, P.K. Development and Characterization of Low-Density Ca-Based Bulk Metallic Glasses: An Overview. Metall. Mater. Trans. A 2008, 39, 1888–1900. [Google Scholar] [CrossRef]
  364. Zhao, K.; Li, J.F.; Zhao, D.Q.; Pan, M.X.; Wang, W.H. Degradable Sr-based bulk metallic glasses. Scr. Mater. 2009, 61, 1091–1094. [Google Scholar] [CrossRef]
  365. Jiao, W.; Zhao, K.; Xi, X.K.; Zhao, D.Q.; Pan, M.X.; Wang, W.H. Zinc-based bulk metallic glasses. J. Non-Cryst. Solids 2010, 356, 1867–1870. [Google Scholar] [CrossRef]
  366. Inoue, A.; Nishiyama, N. New Bulk Metallic Glasses for Applications as Magnetic-Sensing, Chemical, and Structural Materials. MRS Bull. 2007, 32, 651–658. [Google Scholar] [CrossRef]
  367. Kim, S.C.; Yamaura, S.-I.; Shimizu, Y.; Nakashima, K.; Igarashi, T.; Makino, A.; Inoue, A. Production of Ni65Cr15P16B4 metallic glass-coated bipolar plate for fuel cell by high velocity oxy-fuel (HVOF) spray coating method. Mater. Trans. 2010, 51, 1609–1613. [Google Scholar] [CrossRef]
  368. Chen, Z.; Yang, Y.; Kumar, S.; Lu, G. First-Principles Prediction of Oxygen Reduction Activity on Pd−Cu−Si Metallic Glasses. J. Phys. Chem. C 2014, 118, 48. [Google Scholar] [CrossRef]
  369. Tian, R.; Qin, Z. Bulk metallic glass Zr55Cu30Al10Ni5 bipolar plates for proton exchange membrane fuel cell. Energy Convers. Manag. 2014, 86, 927–932. [Google Scholar] [CrossRef]
  370. Jayalakshmi, S.; Vasantha, V.S.; Fleury, E.; Gupta, M. Characteristics of Ni–Nb-based metallic amorphous alloys for hydrogen-related energy applications. Appl. Energy 2012, 90, 94–99. [Google Scholar] [CrossRef]
  371. Tkachov, V.I. Problems of hydrogen degradation of metals. Mater. Sci. 2000, 36, 481–488. [Google Scholar] [CrossRef]
  372. Kirchheim, R. Hydrogen solubility and diffusivity in defective and amorphous metals. Prog. Mater. Sci. 1988, 32, 261–325. [Google Scholar] [CrossRef]
  373. Huang, J.C.; Chu, J.P.; Jang, J.S.C. Recent progress in metallic glasses in Taiwan. Intermetallics 2009, 17, 973–987. [Google Scholar] [CrossRef]
  374. Kim, H.; Gilmore, C.M.; Piqué, A.; Horwitz, J.S.; Mattoussi, H.; Murata, H.; Kafafi, Z.H.; Chrisey, D.B. Electrical, optical, and structural properties of indium–tin–oxide thin films for organic light-emitting devices. J. Appl. Phys. 1999, 86, 6451. [Google Scholar] [CrossRef]
  375. Hartnagel, H.L.; Dawar, A.L.; Jain, A.K. Semiconducting Transparent Thin Films. MRS Bull. 1997, 22, 66–69. [Google Scholar]
  376. Lee, C.J.; Lin, H.K.; Sun, S.Y.; Huang, J.C. Characteristic difference between ITO/ZrCu and ITO/Ag bi-layer films as transparent electrodes deposited on PET substrate. Appl. Surf. Sci. 2010, 257, 239–243. [Google Scholar] [CrossRef]
  377. Wang, W.H.; Hsu, J.H.; Huang, J.C. Optical reflectivity improvement by upgrading metallic glass film quality. Appl. Phys. Lett. 2013, 103, 161906. [Google Scholar] [CrossRef]
  378. Hu, T.T.; Hsu, J.H.; Huang, J.C.; Kuan, S.Y.; Lee, C.J.; Nieh, T.G. Correlation between reflectivity and resistivity in multi-component metallic systems. Appl. Phys. Lett. 2012, 101, 011902. [Google Scholar] [CrossRef]
  379. Seddon, A.B. Chalcogenide glasses: A review of their preparation, properties and applications. J. Non-Cryst. Solids 1995, 184, 44–50. [Google Scholar] [CrossRef]
  380. Burgess, T.; Ferry, M. Nanoindentation of metallic glasses. Mater. Today 2009, 12, 24–32. [Google Scholar] [CrossRef]
  381. Henao, J.; Conctustell, A.; Cano, I.G.; Dosta, S.; Cinca, N.; Guilemany, J.M.; Suhonen, T. Novel Al-based metallic glass coatings by Cold Gas Spray. Mater. Des. 2016, 94, 253–261. [Google Scholar] [CrossRef]
  382. Kim, C.H.; Jang, Y.H.; Hwang, H.J.; Sun, Z.H.; Moon, H.B.; Cho, J.H. Observation of bistable resistance memory switching in CuO thin films. Appl. Phys. Lett. 2009, 94, 102107. [Google Scholar] [CrossRef]
  383. Sun, B.; Liu, Y.X.; Liu, L.F.; Xu, N.; Wang, Y.; Liu, X.Y.; Han, R.Q.; Kang, J.F. Highly uniform resistive switching characteristics of TiN/ZrO2/Pt memory devices. J. Appl. Phys. 2009, 105, 061630. [Google Scholar] [CrossRef]
  384. Hong, S.; Long, D.X.; Hwang, I.; Kim, J.-S.; Park, Y.C.; Kang, S.-O.; Park, B.H. Unipolar resistive switching mechanism speculated from irreversible low resistance state of Cu2O films. Appl. Phys. Lett. 2011, 99, 052105. [Google Scholar] [CrossRef]
  385. Kim, K.M.; Choi, B.J.; Koo, B.W.; Choi, S.; Jeong, D.S.; Hwang, C.S. Resistive switching in Pt/Al2O3/TiO2/Ru stacked structures. Electrochem. Solid-State Lett. 2006, 9, 343–346. [Google Scholar] [CrossRef]
  386. Tulu, B.; Chang, W.Z.; Chu, J.P.; Wang, S.F. Forming-free resistive switching characteristics of 15 nm-thick multicomponent oxide. Appl. Phys. Lett. 2013, 103, 252904. [Google Scholar] [CrossRef]
  387. Wang, X.; Li, X.; Xie, H.; Fan, T.; Zhang, L.; Li, K.; Cao, Y.; Yang, X.; Liu, B.; Bai, P. Effects of Al and La elements on mechanical properties of CoNiFe0.6Cr0.6 high-entropy alloys: A first-principles study. J. Mater. Res. Technol. 2023, 23, 1130–1140. [Google Scholar] [CrossRef]
  388. Zhang, L.T.; Pelletier, J.M.; Qiao, J.C. Dynamic mechanical behavior of (La0.7Ce0.3)65Al10Co25 bulk metallic glass: Influence of the physical aging and heat treatment. J. Alloys Compd. 2023, 869, 159271. [Google Scholar] [CrossRef]
  389. Zhang, C.; Khorshidi, H.; Najafi, E.; Ghasemi, M. Fresh, mechanical and microstructural properties of alkali-activated composites incorporating nanomaterials: A comprehensive review. J. Clean. Prod. 2023, 384, 135390. [Google Scholar] [CrossRef]
Figure 1. The effect of temperature during the solidification of Al86Ni8Y6 is shown [86]. Adding Y2O3 particles into the liquid melt forms an Al23Ni6Y4 intermetallic. Finally, the fcc α-Al solid solidifies at the interface of the Al23Ni6Y4 intermetallic. A time–temperature–transformation (TTT) diagram shows the influence of heating and cooling rates on the crystallization behavior [86].
Figure 1. The effect of temperature during the solidification of Al86Ni8Y6 is shown [86]. Adding Y2O3 particles into the liquid melt forms an Al23Ni6Y4 intermetallic. Finally, the fcc α-Al solid solidifies at the interface of the Al23Ni6Y4 intermetallic. A time–temperature–transformation (TTT) diagram shows the influence of heating and cooling rates on the crystallization behavior [86].
Coatings 13 01689 g001
Figure 2. (ac) Schematic drawing of the experimental configuration of an ultrafast liquid-quenching process [87]. (a) The process begins by bringing two protruded nano-tips in contact, which are melted by the application of an electric pulse; (b) heat is dissipated through the bulk substrates, and (c) the melting zone transforms into monatomic MGs due to liquid quenching. (d) High-resolution transmission electron microscope (HRTEM) image showing two contacting Ta nano-tips before the application of the electric pulse. The boundary of the two-contacting nano-tips are shown using the dotted lines (e) HRTEM image showing vitrified Ta MG after the application of the electric pulse. (fh) Fast Fourier transformations confirming a fully vitrified region in the middle (g) bounded by two crystalline substrates.
Figure 2. (ac) Schematic drawing of the experimental configuration of an ultrafast liquid-quenching process [87]. (a) The process begins by bringing two protruded nano-tips in contact, which are melted by the application of an electric pulse; (b) heat is dissipated through the bulk substrates, and (c) the melting zone transforms into monatomic MGs due to liquid quenching. (d) High-resolution transmission electron microscope (HRTEM) image showing two contacting Ta nano-tips before the application of the electric pulse. The boundary of the two-contacting nano-tips are shown using the dotted lines (e) HRTEM image showing vitrified Ta MG after the application of the electric pulse. (fh) Fast Fourier transformations confirming a fully vitrified region in the middle (g) bounded by two crystalline substrates.
Coatings 13 01689 g002
Figure 3. Scanning electron microscope (SEM) images of (a) the microstructure of the cross section of the Fe 34 Co 34 B 20 Si 5 C 3 Nb 4 amorphous composite coating, (b) that near the substrate/coating interface, (c) a magnified view of rectangular region A, and (d) a magnified view of rectangular region B. The inset in (c) lists the electron probe microanalysis (EPMA) data of the white particle [88].
Figure 3. Scanning electron microscope (SEM) images of (a) the microstructure of the cross section of the Fe 34 Co 34 B 20 Si 5 C 3 Nb 4 amorphous composite coating, (b) that near the substrate/coating interface, (c) a magnified view of rectangular region A, and (d) a magnified view of rectangular region B. The inset in (c) lists the electron probe microanalysis (EPMA) data of the white particle [88].
Coatings 13 01689 g003
Figure 4. Cross-sectional optical microscope images showing (a) atomized powders on a substrate. Samples prepared using selective laser melting (SLM) with (b) v = 2500 mm/s and P = 300 W, (c) v = 2500 mm/s and P = 340 W, and (d) v = 1500 mm/s and P = 340W. (e) Side- and (f) top-view of the final SLM samples. The observed micro-pores upon using v = 2500 mm/s and P = 300 W are shown with white-arrows in (b). [129].
Figure 4. Cross-sectional optical microscope images showing (a) atomized powders on a substrate. Samples prepared using selective laser melting (SLM) with (b) v = 2500 mm/s and P = 300 W, (c) v = 2500 mm/s and P = 340 W, and (d) v = 1500 mm/s and P = 340W. (e) Side- and (f) top-view of the final SLM samples. The observed micro-pores upon using v = 2500 mm/s and P = 300 W are shown with white-arrows in (b). [129].
Coatings 13 01689 g004
Figure 5. Illustration of (a) the crystalline Zr–powder, (b) amorphous CuZrAl powder, and (c) a mixture of 20% vol of Zr-powder with amorphous powder produced using the ball-milling technique [152].
Figure 5. Illustration of (a) the crystalline Zr–powder, (b) amorphous CuZrAl powder, and (c) a mixture of 20% vol of Zr-powder with amorphous powder produced using the ball-milling technique [152].
Coatings 13 01689 g005
Figure 6. (a) Temperature and pressure cycle of a spark plasma sintering (SPS) densification technique, and (b) cylindrical sample produced from the powder after densification [152].
Figure 6. (a) Temperature and pressure cycle of a spark plasma sintering (SPS) densification technique, and (b) cylindrical sample produced from the powder after densification [152].
Coatings 13 01689 g006
Figure 7. (a) Schematic illustration of magnetron co-sputtering of the Zr-Cu-Al-Ag system, (b) deposited films on the Si wafer where the compositional variation of the elements are obtained by tilting the guns at different angles, and (c) antibacterial activity analysis of the deposited films [168].
Figure 7. (a) Schematic illustration of magnetron co-sputtering of the Zr-Cu-Al-Ag system, (b) deposited films on the Si wafer where the compositional variation of the elements are obtained by tilting the guns at different angles, and (c) antibacterial activity analysis of the deposited films [168].
Coatings 13 01689 g007
Figure 8. X-ray diffraction (XRD) and high-resolution transmission electron (HR-TEM) microscopy (a,f,k) images of Zr-Ti, Fe-Al and Zr-TI-Fe-Al systems, where the insets (b,g,l) show selected-area electron diffractions (SAED). The corresponding fast Fourier transformation (FFT), (c,h,m), inverse FFT (d,i,n), and filtered FFT (e,j,o) images of the white boxed region are also shown [198].
Figure 8. X-ray diffraction (XRD) and high-resolution transmission electron (HR-TEM) microscopy (a,f,k) images of Zr-Ti, Fe-Al and Zr-TI-Fe-Al systems, where the insets (b,g,l) show selected-area electron diffractions (SAED). The corresponding fast Fourier transformation (FFT), (c,h,m), inverse FFT (d,i,n), and filtered FFT (e,j,o) images of the white boxed region are also shown [198].
Coatings 13 01689 g008
Figure 9. (a) Strength vs. ductility relationships of Cu-Zr-based metallic glass compared to conventional engineering alloys. (b) Tensile properties of different metallic-glass systems. The ductility and yield strength of CuZr—based metallic glass composites is compared with existing literatures [258,260,261,262,263].
Figure 9. (a) Strength vs. ductility relationships of Cu-Zr-based metallic glass compared to conventional engineering alloys. (b) Tensile properties of different metallic-glass systems. The ductility and yield strength of CuZr—based metallic glass composites is compared with existing literatures [258,260,261,262,263].
Coatings 13 01689 g009
Figure 10. Electrochemical analysis of pure aluminum (Al) and different Al-based amorphous systems ( Al 86 Ce 10 X 4 , where X = Ti, Cr, Mn, Fe, Co, Ni and Cu) in NaCl (3.56 wt.%) [303].
Figure 10. Electrochemical analysis of pure aluminum (Al) and different Al-based amorphous systems ( Al 86 Ce 10 X 4 , where X = Ti, Cr, Mn, Fe, Co, Ni and Cu) in NaCl (3.56 wt.%) [303].
Coatings 13 01689 g010
Figure 11. Soft magnetic features of Fe-based ( Fe 80 ( Nb 1 x Mo x ) 5 B 15 ; where, x = 0, 0.15, 0.30, 0.45, 0.60, 0.75) metallic glass. (a) M-H hysteresis loops indicating the relationship between saturation magnetization (Ms) and coercivity (Hc), and (b) the variation in Ms and Hc with the Mo content [313].
Figure 11. Soft magnetic features of Fe-based ( Fe 80 ( Nb 1 x Mo x ) 5 B 15 ; where, x = 0, 0.15, 0.30, 0.45, 0.60, 0.75) metallic glass. (a) M-H hysteresis loops indicating the relationship between saturation magnetization (Ms) and coercivity (Hc), and (b) the variation in Ms and Hc with the Mo content [313].
Coatings 13 01689 g011
Figure 12. Antibacterial activity of Zr−Cu−Al−Ag metallic-glass systems at 9 h. (a) Contour plots of optical density (OD) of all the metallic-glass systems. (b,c) OD vs. time curves representing the highest (red) and lowest (blue) antibacterial activity of the Zr 38 Cu 36 Al 18 Ag 8 and Zr 23 Cu 36 Al 29 Ag 12 systems, respectively, where the controls are shown by the grey lines. The SEM morphologies of E. coli bacteria on the surface of Zr 38 Cu 36 Al 18 Ag 8 (bottom left) and Zr 23 Cu 36 Al 29 Ag 12 (bottom right) systems, respectively [168].
Figure 12. Antibacterial activity of Zr−Cu−Al−Ag metallic-glass systems at 9 h. (a) Contour plots of optical density (OD) of all the metallic-glass systems. (b,c) OD vs. time curves representing the highest (red) and lowest (blue) antibacterial activity of the Zr 38 Cu 36 Al 18 Ag 8 and Zr 23 Cu 36 Al 29 Ag 12 systems, respectively, where the controls are shown by the grey lines. The SEM morphologies of E. coli bacteria on the surface of Zr 38 Cu 36 Al 18 Ag 8 (bottom left) and Zr 23 Cu 36 Al 29 Ag 12 (bottom right) systems, respectively [168].
Coatings 13 01689 g012
Figure 13. Metabolic activity of (a) HAECs and (b) HASMCs at day 1, 3 and 5, where (*) represents statistically a significant difference (p < 0.05; n = 3). SEM images depicting cell morphology at 1, 3, and 5 days for (c) HAECs and (d) HASCMs. (Scale bars = 500 μm and 5 μm for the larger and inset images, respectively) [297].
Figure 13. Metabolic activity of (a) HAECs and (b) HASMCs at day 1, 3 and 5, where (*) represents statistically a significant difference (p < 0.05; n = 3). SEM images depicting cell morphology at 1, 3, and 5 days for (c) HAECs and (d) HASCMs. (Scale bars = 500 μm and 5 μm for the larger and inset images, respectively) [297].
Coatings 13 01689 g013
Figure 14. (a) Unipolar current (I)—voltage (V) curve of a Pt/TFMGO/Pt memory device. (b) Endurance and (c) retention features measured at a reading voltage of 0.1 V for the Pt/TFMGO/Pt memory cell and for a resistance with respect to the switching cycle and time, respectively [386].
Figure 14. (a) Unipolar current (I)—voltage (V) curve of a Pt/TFMGO/Pt memory device. (b) Endurance and (c) retention features measured at a reading voltage of 0.1 V for the Pt/TFMGO/Pt memory cell and for a resistance with respect to the switching cycle and time, respectively [386].
Coatings 13 01689 g014
Figure 15. Synthesis techniques, features, limitations, and potential applications of metallic-glass systems.
Figure 15. Synthesis techniques, features, limitations, and potential applications of metallic-glass systems.
Coatings 13 01689 g015
Table 1. Metallic-glass systems and corresponding parameters associated with Inoue’s empirical glass-forming rules [5,48]. The heat of mixing values was obtained from refs. [5,48] and the atomic size mismatch was calculated by using the atomic radius [54]. The MG fabrication technique is also listed.
Table 1. Metallic-glass systems and corresponding parameters associated with Inoue’s empirical glass-forming rules [5,48]. The heat of mixing values was obtained from refs. [5,48] and the atomic size mismatch was calculated by using the atomic radius [54]. The MG fabrication technique is also listed.
Metallic Glass SystemAtomic Size Mismatch (%)
(rbase − rO)/rbase
Heat of Mixing
(kJ/mol)
Fabrication RouteRef.
Zr41.2Ti13.8Cu12.5Ni10Be22.5Zr:Ti = 9.6; Zr:Cu = 20
Zr:Ni = 22.3
Zr:Be = 29.6
Zr-Ti = 0; Zr-Cu = −23
Zr-Ni = −49; Zr-Be = −43
Ti-Cu = −9; Ti-Ni = −35
Ti-Be = −30; Cu-Ni = 4
Cu-Be = 0; Ni-Be = −4
Casting in copper molds[11]
Pt57.5Cu14.7Ni5.3P22.5Pt:Cu = 7.8
Pt:Ni = 10.2
Pt:P = 23.5
Pt-Cu = −12; Pt-Ni = −5
Pt-P = −34.5; Cu-Ni = 4
Cu-P = −17.5; Ni-P = −34.5
Water quenching[55]
Zr-Ti-Nb-Cu-BeZr:Ti = 9.6
Zr:Nb = 10.7
Zr:Cu = 20
Zr:Be = 29.6
Zr-Ti = 0; Zr-Nb = 4
Zr-Cu = −23; Zr-Be = −43
Ti-Nb = 2; Ti-Cu = −9
Ti-Be = −30; Nb-Cu = 3
Nb-Be = −25; Cu-Be = 0
Arc-melting and heated via induction[56]
Cu47.5Zr47.5Al5Cu:Zr = 20
Cu:Al = 10.7
Zr:Al = 10.6
Cu-Zr = −23; Cu-Al = −1
Zr-Al = −44
Arc-melting[57]
Pd40Ni40P20Pd:Ni = 9.4
Pd:P = 23
Ni:P = 14.8
Pd-Ni = 0; Pd-P = −36.5
Ni-P = −34.5
Fluxing[8]
Zr39Cu39Ag22Zr:Cu = 20
Zr:Ag = 9.8
Zr-Cu = −23; Zr-Ag = −20; Cu-Ag = 2DC reactive magnetron sputtering[58]
Zr59Ti22Ag19Zr:Ti = 9.6
Zr:Ag = 9.8
Zr-Ti = 0; Zr-Ag = −69; Ti-Ag = −2Magnetron sputtering[59]
Zr-Ti-FeZr:Ti = 9.6
Zr:Fe = 22.5
Ti:Fe = 8.8
Zr-Ti = 0; Zr-Fe = −25;
Ti-Fe = −17
Magnetron co-sputtering[60]
Zr-Ni-Al-SiZr:Ni = 22.3
Zr:Al = 10.6
Zr:Si = 28
Zr-Ni = −49; Zr-Al = −44; Zr-Si = −84; Ni-Al = −22; Ni-Si = −40; Al-Si = −19RF and DC reactive magnetron sputtering[61]
Zr-Cu-Al-AgZr:Cu = 20
Zr:Al = 10.6
Zr:Ag = 9.8
Zr-Cu = −23; Zr-Al = −44
Zr-Ag = −69; Cu-Al = −1
Cu-Ag = 2; Al-Ag = −4
DC magnetron sputtering[62]
Cu48Zr42Ti4Al6Cu:Zr = 20
Cu:Ti = 14.3
Cu:Al = 10.7
Cu-Zr = −23; Cu-Ti = −9
Cu-Al = −1; Zr-Ti = 0
Zr-Al = −44; Ti-Al = −30
RF magnetron sputtering[63]
Zr60.14Cu22.31Fe4.85Al9.7Ag3Zr:Cu = 20
Zr:Fe = 22.5
Zr:Al = 10.6
Zr:Ag = 9.8
Zr-Cu = −23; Zr-Fe = −25
Zr-Al = −44; Zr-Ag = −69
Cu-Fe = 13; Cu-Al = −1 Cu-Ag = 2; Fe-Al = −11
Fe-Ag = 28; Al-Ag = −4
DC Magnetron sputtering[64]
Zr59Cu20Al10Ni8Ti3Zr:Cu = 20
Zr:Al = 10.6
Zr:Ni = 22.3
Zr:Ti = 9.6
Zr-Cu = −23; Zr-Al = −44
Zr-Ni = −49; Zr-Ti = 0
Cu-Al = −1; Cu-Ni = 4
Cu-Ti = −9; Al-Ni = −22;
Ni-Ti = −35
Arc-melting[65]
Fe40Ni40P14B6Fe:Ni = 0.3
Fe:P = 14.5
Fe:B = 34.1
Ni:P = 14.8
Ni:B = 34.1
Fe-Ni = −2; Fe-P = −39.5
Fe-B = −26; Ni-P = −34.5
Ni-B = −24; P-B = 0.5
Induction melting, fluxing, re-melt, and quenching[66]
Fe50.26B2.62Si2.41Cr23.86Mo20.85Fe:B = 34.1
Fe:Si = 7.1
Fe:Cr = 0.6
Fe:Mo = 9.8
Fe-B = −26; Fe-Si = −35
Fe-Cr = −1; Fe-Mo = −2
B-Si = −14; B-Cr = −31
B-Mo = −7; Si-Cr = −37
Si-Mo = −35; Cr-Mo = 0
Atmospherically plasma-sprayed[67]
Ti–Ni–Cu–Sn, Ti–Ni–Cu–Sn–Be and Ti–Ni–Cu–Sn–Be–ZrTi:Ni = 14.8
Ti:Cu = 12.6
Ti-Sn = 8.1
Ti:Be = 22.8
Ti:Zr = 9.6
Ti-Ni = −35; Ti-Cu = −9
Ti-Sn = −21; Ti-Be = −30;
Ti-Zr = 0; Ni-Cu = 4
Ni-Sn = −4; Ni-Be = −4
Ni-Zr = −49; Cu-Sn = 7
Cu-Be = 0; Cu-Zr = −23
Sn-Be = 15; Sn-Zr = −43
Be-Zr = −43
Injection casting[68]
Cu47Ti34Zr11Ni8, Cu47Ti33Zr11Ni8Fe1 and Cu47Ti33Zr11Ni8Si1Cu:Ti = 14.3
Cu:Zr = 25.4
Cu:Ni = 2.6
Cu:Fe = 2.9
Cu:Si = 9.8
Cu-Ti = −9; Cu-Zr = −23
Cu-Ni = 4; Cu-Fe = 13
Cu-Si = −19; Ti-Zr = 0
Ti-Ni = −35; Ti-Fe = −17
Ti-Si = −66; Zr-Ni = −49
Zr-Fe = −25; Zr-Si = −84
Ni-Fe = −2; Ni-Si = −40
Copper mold casting[69]
Table 2. Metallic-glass systems are fabricated by different solidification techniques and associated cooling rates.
Table 2. Metallic-glass systems are fabricated by different solidification techniques and associated cooling rates.
Solidification TechniqueCooling Rate
(K/s)
Fabricated SystemRef.
Conventional die casting101–103Zr46.75Ti8.25Cu7.5Ni10Be27.5, Zr–Al–Cu, La55Al25Ni20, Mg80Cu10Y10,[6,12,53,73]
Melt spinning105–106Fe57.2Co30.8Zr7−xHfxB4Cu1 (x = 3, 5, and 7), Zr61Cu17.5Ni10Al7.5Si4[74,75,76]
Liquid splat-quenching∼109–1010Zr46.7Ti8.3Cu7.5Ni10Be27.5 Au-Si, Au0.778Ge0.138Si0.084[2,77,78,79]
Pulsed laser quenching∼1012–1013Cu-Ti-Zr, Cu-Ti, NixNb100−x[80,81,82,83]
Nano calorimetry104–106Au–Cu–Si[84,85]
Table 3. Metallic-glass systems joined via different welding techniques.
Table 3. Metallic-glass systems joined via different welding techniques.
Liquid-state weldingMetallic Glass SystemWelded toWelding TechniqueParametersRef.
Thickness (mm)Power (kW)Scanning Speed (mm/s)
Zr41Be23Ti14Cu12Ni10Zr55Al10Ni5Cu30Electron-beam3.5933[104]
Zr41Ti14Cu12Ni10Be23Polycrystalline Zr metalElectron-beam3933[105]
Zr41Be23Ti14Cu12Ni10Ti metalElectron-beam3966[106]
Zr41Be23Ti14Cu12Ni10Stainless steelElectron-beam2966[107]
Zr45Cu48Al7Zr45Cu48Al7Laser11.233–133[90]
Cu54Ni6Zr22Ti18Cu54Ni6Zr22Ti18Pulsed Nd:YAG61.50.33 and 1[108]
Zr55Al10Ni5Cu30304 austenitic stainless steelFiber-laser92–101.2[109]
(Zr53Cu30Ni9Al8)Si0.5(Zr53Cu30Ni9Al8)Si0.5Pulsed Nd:YAG11.3–1.71[110]
Pd43Cu27Ni10P20Pd43Cu27Ni10P20Pulsed laser beam10.750–1.1250.33[111]
Solid-state weldingMetallic glass systemWelded toWelding processParameters
Rotational speed (rpm)Time (s)
Ti40Zr25Ni3Cu12Be20Ti40Zr25Ni3Cu12Be20Friction1800–22005–7[103]
Zr41.5Ti13.8Cu12.5Ni10Be22.5Zr41.5Ti13.8Cu12.5Ni10Be22.5Friction27000–35[112]
Zr55Al10Ni5Cu30Zr55Al10Ni5Cu30Friction18000.4–1.0[113]
Pd40Ni40P20Pd40Cu30Ni10P20Friction60000.2[114]
Table 4. Metallic-glass systems produced via additive manufacturing process.
Table 4. Metallic-glass systems produced via additive manufacturing process.
Materials SystemsLaser Power (W)Scanning Speed
(mm/s)
Hatch Spacing
(μm)
Ref.
Zr55Cu30Ni5Al102401200100[120]
Al86Ni6Y4.5Co2La1.5120750100[132]
Al85N5Y6Co2Fe2200625150[131]
Ti47Cu38Zr7.5Fe2.5Sn2Si1Ag2602000140[133]
Zr52.5Cu17.9Ni14.6Al10Ti530–120250–2000100–200[134]
Fe74Mo4P10C7.5B2.5Si23203470124[123]
Zr52.5Ti5Cu17.9Ni14.6Al10200500150[130]
Fe43.7Co7.3Cr14.7Mo12.6C15.5B4.3Y1.9150–350200–1000-[135]
Fe54.35Cr18.47Mn2.05Mo13.93W5.77B3.22C0.90Si1.32220–380200090[136]
Zr50Ti5Cu27Ni10Al820013.3-[137]
Table 5. Metallic-glass systems produced via powder metallurgy.
Table 5. Metallic-glass systems produced via powder metallurgy.
Name of the SystemPowder Production TechniqueDensification ProcessRef.
Ni59Zr20Ti16Si2Sn3Argon gas atomizationWarm extrusion[143]
Cu50Ti32Zr12Ni5Si1High-pressure gas-atomizationEqual channel angular extrusion[144]
Mg65Cu25Gd10Mechanically milledSpark plasma sintering[138]
Ni53Nb20Ti10Zr8Co6Ta3Mechanical alloyingSpark plasma sintering[139]
Al82La10Ni4Fe4,Mechanical alloyingSpark plasma sintering[158]
Ni52.5Nb10Zr15Ti15Pt7.5Argon gas atomizationSpark plasma sintering[159]
Zr55Cu30Al10Ni5Argon gas atomizationSpark plasma sintering[155]
Ni59Zr15Ti13Si3Sn2Nb7Al1Gas-atomizationSpark plasma sintering[160]
(Fe0.72B0.24Nb0.04)95.5Y4.5Gas-atomizationSpark plasma sintering[161]
Fe73Si7B17Nb3Argon gas atomizationSpark plasma sintering[162]
Fe48Cr15Mo14Y2C15B6Argon gas atomizationSpark plasma sintering[163]
Fe67Co9.5Nd3Dy0.5B20Mechanically rotor-milledSpark plasma sintering[164]
Ti50Cu23Ni20Sn7Mechanically milledSpark plasma sintering[165]
Table 6. Parameters associated with the magnetron-sputtering technique used to fabricate various metallic-glass systems.
Table 6. Parameters associated with the magnetron-sputtering technique used to fabricate various metallic-glass systems.
SystemSputtering TypeParametersRef.
DC Power (W)RF Power (W)Base Pressure (Pa)Working Pressure (Pa)Target-Substrate Distance
(mm)
Sputtering Rates (nm/min)Thickness (µm)
Zr–Ni–Al–SiReactive magnetron co-sputteringZr: 700
Ni: 120
Al: 240
Si: 300
2 × 10−40.42502.1–6.1-[61]
Zr61All7.5Ni10Cu17.5Si4DC plus magnetron sputtering30-7.9 × 10−40.5360-0.5[170]
ZrCuAlRF magnetron sputtering-Zr: 225
Cu: 50
Al: 26
6.6 × 10−51.3165-2[171]
Zr55Cu31Ti14DC co-sputtering300-6.6 × 10−5---2–3[172]
Zr-PdRF magnetron sputtering-702–4 × 10−42706.6-[173]
ZrCuAlAgDC magnetron sputtering15–30--0.4607.6–14.50.2[62]
Zr47Cu31Al13Ni9RF magnetron sputtering-100-0.27--0.2[174]
Al48Ag37Ti15Magnetron sputtering--6.6 × 10−50.4--0.5[59]
Ta-Ti-Zr-SiDC co-sputtering--6.6 × 10−5---0.6[175]
Zr60.14Cu22.31Fe4.85Al9.7Ag3DC magnetron sputtering120-2 × 10−40.65--0.53[64]
Table 7. Mechanical properties of various metallic-glass systems. U, C, Y, and F indicate ultimate, compressive, yield and fracture strengths, respectively. For hardness values, V and C indicate Vicker’s and compressive hardness, respectively.
Table 7. Mechanical properties of various metallic-glass systems. U, C, Y, and F indicate ultimate, compressive, yield and fracture strengths, respectively. For hardness values, V and C indicate Vicker’s and compressive hardness, respectively.
Name of the SystemsStrength
(GPa)
Elastic Modulus
(GPa)
Hardness
(GPa)
Measurement TechniqueRef.
Zr-based MGsZr41.25Ti13.75Cu12.5Ni10Be22.51.9 (U)965.23 (V)Tensile and compression test[223]
(Zr55Al10Ni5Cu30)98.5Si1.51.8 (U)875.2 (V)Tensile and compression test[224]
Zr61Cu17.5Ni10Al7.5Si41.8 (C)--5 (C)Compressive strength and hardness[75]
Zr46Cu37.6Ag8.4Al81.9 (Y)925.4 (V)Vickers microhardness[216]
Zr40Ti14Ni10Cu12Be242.3 (Y)1149.7Nanoindentation[225]
ZrTiCuNiAl1.96 (Y)--5.5Nanoindentation[226]
Zr52.5Al10Ti5Cu17.9Ni14.60.82 (Y)109--Nanoindentation[227]
Zr69.5Cu12Ni11Al7.5--93.865.66Nanoindentation[228]
ZrTiAlFeCuAg (Zr = 60%)1.58 ± 0.03 (Y)78 ± 14.5 ± 0.06 (V)Compression, notch-toughness tests, and ultrasound spectroscopy[214]
Ti-based MGs(Ti40Zr10Cu38Pd12)100–x Nbx (x = 0, 2, 3, 4)1.2–2 (F)100–1066–8 (V)Compression test[229]
Ti40Zr25Ni3Cu12Be20 (10nm/2.50mN/s)----8.29 ± 0.13Nanoindentation[230]
Ti75Zr10Si152.6 (Y)--0.007Microhardness[54]
Ti60Nb15Zr10Si152.2 (Y)--0.006Microhardness[54]
Ti40Cu36Pd14Zr10--1107.7Microscratch[231]
Ti50Cu25Ni15Sn3Be72.17 (C)--6.5Uniaxial compression[232]
Ti47Cu38Zr7.5Fe2.5Sn2Si1Ag22.08 (C)100.4 ± 0.15.7 ± 0.05Compression[178]
Cu-based MGsCu60Zr30Ti101.78 (Y)
2 (F)
1126.4 (V)Tensile and compression deformation[233]
Cu60Hf25Ti151.92 (Y)
2.13 (F)
1206.6 (V)Tensile and compression deformation[233]
Cu60Zr30Ti10--93.88 ± 1.77.61 ± 0.33Nanoindentation[234]
(Cu0.6Hf0.25Ti0.15)90Nb102.073 (Y)
2.232 (F)
106--Compression test[235]
Cu47Ti33Zr12Ni8Si12.087 (F)118.6--Compression test[236]
(Cu50Zr50)90Al10--117.35.3 (V)
8.7
Microhardness
nanoindentation
[237]
Cu49Hf42Al92.408 (Y)
2.620 (F)
102--Compression test[238]
(Cu0.50Hf0.35Ti0.10Ag0.05)97Ta32.510 (F)151.26.04Compression test
nanoindentation
[239]
Fe-based MGFe59Cr6Mo14C15B63.8 (Y)
4.4 (F)
204∼11Compression tests[222]
Fe41Co7Cr15Mo14Y2C15B63.5 (F)26512.3 (V)Compression, bending and hardness tests[240]
Fe36Co36B19.2Si4.8Nb4>4 (Y)201 ± 10
192 ± 0.5
14Compression test
nanoindentation
[241]
(Fe0.75B0.15Si0.1)96Nb3.25 (Y)17510.4 (V)Compression Test[242]
[(Fe0.8Co0.2)0.75B0.2Si0.05]96Nb44.05 (Y)
4.17 (F)
20512.01 (V)Vickers hardness compression test[243]
Fe66Mo10P12C10B22.55 (Y)
3.25 (F)
1768.83 (V)Microhardness, compression and resonant ultrasound spectroscopy test[244]
Fe0.432Co0.288B0.192Si0.048Nb0.04)98Cr2∼4.0 (Y)--11.51–12.51Nanoindentation[245]
Pd,Ni,Mg,Ca,Al-Based MGsPd40Ni40P201.78 (Y)103–108--Nanoindentation[246]
Ni60Nb37Sn33.7--8.83 (V)Vickers hardness measurements[247]
Mg66Zn29Ca4Ag1----2.35 ± 0.03Microhardness[248]
Ca20Mg20Zn20Sr20Yb200.37 ± 0.025 (F)19.4 ± 3.4--Uniaxial compression test[249]
Al85Y10Ni50.9262.83.7Microhardness and tensile[250]
Mg65Cu25Gd10∼0.8 (C)--∼2.5Compression test[75]
Table 8. Magnetic properties of various Fe- and Co-based metallic-glass systems.
Table 8. Magnetic properties of various Fe- and Co-based metallic-glass systems.
Name of the SystemsCoercivity
(A/m)
Saturation Magnetization (T)PermeabilityCurrie Temperature
(K)
Ref.
Fe-basedFe70Al5Ga2P9.65C5.75B4.6Si32.21.2110,000620[318]
[(Fe1−xCox)75B20Si5]93Nb4Y3 Ribbon (x = 0, 0.2, 0.4, 0.6)2–15∼0.4–0.6----[319]
Fe72Al5Ga2P11C6B45.11.079000596–605[320]
[(Fe1−xCox)0.75B0.2Si0.05]96Nb41.5–2.70.84–1.13>1.2 × 104600–690[321]
Fe76Si9B10P50.81.51----[322]
Fe76−xC7.0-Si3.3B5.0P8.7Cux (x = 0, 0.3, 0.7, 1.0 at.%)111.79----[323]
Fe66Co10Mo3.5P10C4B4Si2.51.01.23450,000--[324]
Fe82.75Si4B8P4Cu1.252.11.8331,600--[325]
Fe-Si-B-M
(M = Cu, Nb, Mo, W, Ta)
6.91.416000631[326]
Fe55Co30Cu1Nb7Si1B85.91.711150--[327]
Co-basedCo66Fe4Mo2Si16B1----∼109,000--[328]
CobalFe4Ni2Si15B14<1–2--24,000490[329]
Co33.9Fe33.9B22.5Si5.7Nb44.90.98----[330]
Co40Fe27Zr3Ti3Mo1.5Si1.5B24
(Cylinder)
81.2----[331]
Co42Fe20Hf3Mo3Ti3B2920.6----[332]
Co67Fe4Mo2Si17B11 (annealed at 360°)∼0.1∼1.2----[333]
Co68.15Fe4.35Si12.5B152100.81----[334]
Co43Fe20Ta5.5B31.50.250.49550,000--[335]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Jabed, A.; Bhuiyan, M.N.; Haider, W.; Shabib, I. Distinctive Features and Fabrication Routes of Metallic-Glass Systems Designed for Different Engineering Applications: A Review. Coatings 2023, 13, 1689. https://doi.org/10.3390/coatings13101689

AMA Style

Jabed A, Bhuiyan MN, Haider W, Shabib I. Distinctive Features and Fabrication Routes of Metallic-Glass Systems Designed for Different Engineering Applications: A Review. Coatings. 2023; 13(10):1689. https://doi.org/10.3390/coatings13101689

Chicago/Turabian Style

Jabed, Akib, M. Nabil Bhuiyan, Waseem Haider, and Ishraq Shabib. 2023. "Distinctive Features and Fabrication Routes of Metallic-Glass Systems Designed for Different Engineering Applications: A Review" Coatings 13, no. 10: 1689. https://doi.org/10.3390/coatings13101689

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