Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 4390 2022-11-24 16:54:36 |
2 Reference format revised. Meta information modification 4390 2022-11-28 03:15:12 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Biały, M.;  Hasiak, M.;  Łaszcz, A. Prospect Biomedical Applications of Novel Functional Metallic Glasses. Encyclopedia. Available online: https://encyclopedia.pub/entry/36632 (accessed on 27 July 2024).
Biały M,  Hasiak M,  Łaszcz A. Prospect Biomedical Applications of Novel Functional Metallic Glasses. Encyclopedia. Available at: https://encyclopedia.pub/entry/36632. Accessed July 27, 2024.
Biały, Michał, Mariusz Hasiak, Amadeusz Łaszcz. "Prospect Biomedical Applications of Novel Functional Metallic Glasses" Encyclopedia, https://encyclopedia.pub/entry/36632 (accessed July 27, 2024).
Biały, M.,  Hasiak, M., & Łaszcz, A. (2022, November 25). Prospect Biomedical Applications of Novel Functional Metallic Glasses. In Encyclopedia. https://encyclopedia.pub/entry/36632
Biały, Michał, et al. "Prospect Biomedical Applications of Novel Functional Metallic Glasses." Encyclopedia. Web. 25 November, 2022.
Prospect Biomedical Applications of Novel Functional Metallic Glasses
Edit

The continuous development of novel materials for biomedical applications is resulting in an increasingly better prognosis for patients. The application of more advanced materials relates to fewer complications and a desirable higher percentage of successful treatments. New, innovative materials being considered for biomedical applications are metallic alloys with an amorphous internal structure called metallic glasses. They are currently in a dynamic phase of development both in terms of formulating new chemical compositions and testing their properties in terms of intended biocompatibility. 

metallic glass bulk metallic glass biocompatibility biomedical mechanical properties corrosion resistance

1. Introduction

One of the widely used standard materials groups for medical implants, surgical tools, and other bio-related devices are metallic alloys, including stainless steels, pure Ti and Ti-based alloys, Co-based alloys, pure Zr and Zr-based alloys, bioresorbable Mg-based alloys, pure Ta, and other miscellaneous [1][2][3]. However, they are often afflicted with many problems including insufficient corrosion fatigue, fretting fatigue resistance, as well as possible toxicity by the release of toxic ions. For the stainless steels and Mg-based materials, corrosion resistance is also problematic. For the stainless steels and Ti-based materials, the wear resistance is a serious concern. Moreover, in the case of implants, the stress shielding effect [4]—which consists of transferring the loads through the stiffer element—is challenging. In the typical case, the load is transferred through the implant (stiffer than the bone) without stimulation of surrounding bones to maintain their properties, which causes bone resorption and implant loosening. Materials such as stainless steels, Co-based materials, pure Ta, and certain alloys from other groups—which are characterized by a Young’s modulus significantly higher than the value for the bones (5.7–18.2 GPa [5])—are particularly prone to triggering this effect. However, it was reported that materials with Young’s modulus values of about 60 GPa are already effective in eliminating the stress shielding effect [6]. Ultimately, the mentioned issues lower the applicational biocompatibility of current materials for biomedical applications. Other problems relate to the inapplicability of some materials for magnetic resonance due to the displacement possibility (e.g., for ferromagnetic materials such as ferritic and martensitic stainless steels [7]) or inducing severe artifacts (e.g., for austenitic stainless steels, Co-based alloys, and even Ti-based alloys [8]). Similarly, artifacts can be induced during X-ray imaging [9].

The majority of metallic alloys at room temperature are characterized by a crystalline structure, which ensures the optimal arrangement of atoms [10]. However, to some extent, it is possible to control the structure at the manufacturing stage by altering the thermodynamic conditions. The structures obtained in rapid cooling processes were noticed in metals as early as in the 1940s [11][12] while applying metal vapors to the cold substrate. The first confirmed formation of an amorphous structure in metals obtained by cooling them from liquid to solid state dates to 1960 [13]. It was produced at that time by the Duwez research team at Caltech through the very rapid cooling of an Au-Si alloy into the shape of thin foil. This new group of metallic materials was named metallic glasses, by similarity to well-known amorphous oxide glasses.

The confirmation of the possibility of changing the internal structure in metallic alloys initiated rapid progress and research on this material group. This significant development resulted in the discovery of many new compositions of alloys with higher Glass Forming Ability (GFA), the determination of changes in physical and mechanical properties in MGs, and the development of new production methods. Firstly, the MGs were fabricated only in the form of thin ribbons or foils due to the necessity of preserving the critical cooling rate (≥106 K/s) essential to obtain the amorphous internal structure. Later, the optimization of chemical compositions resulted in the opportunity for manufacturing MGs with lower critical cooling rates (≤103 K/s) and greater critical dimensions exceeding 1 mm, so called Bulk Metallic Glasses (BMGs) [14][15], which now can be produced with critical dimensions measured even in centimeters [16][17]. Eventually, since their discovery, many MGs and BMGs compositions based on various elements have been developed. However, the necessity to maintain the critical cooling rate still restricts the maximal, composition-dependent, achievable dimensions.

The non-crystalline internal structure of metallic glasses implies far-reaching changes in physicochemical properties in relation to their crystalline counterparts. This includes most of the crucial mechanical [18][19][20], electrical [21], magnetic [19][20][22] or corrosion parameters [23]. Typically, the results are, among others, increased strength, hardness, toughness, and wear resistance with a lowered Young’s modulus and low plasticity [24][25]. In terms of strength, the MGs are closer to the maximum theoretical value than any other known metallic material [26]. Additionally, they are characterized by increased corrosion resistance [27][28][29][30], as a result of grain boundary absence that reduces the possible corrosion routes and the common presence of alloying elements with the great ability to chemical passivation. This increase is also connected with nearly ideal chemical and structural (including lack of crystalline inclusions) composition homogeneity in the microscale [31], which reduces the number of galvanic micro-cells formed [32]
With the unique, different than crystalline metallic materials, properties, the MGs and BMGs were already applied or are under thorough consideration for many prospective applications, including biomedical purposes, to form the basis for the next generation of modern devices. In recent times, several metallic glasses were developed specifically for bio-related applications.

2. Zr-Based Metallic Glasses

Some of the most promising metallic glasses for biomedical applications are materials based on zirconium, which is characterized by the lack of local and systemic toxicity as well as the formation of stable oxide layers on the surface in biological environments [33]. Moreover, metallic glasses containing zirconium are among the best alloys in terms of glass forming ability [34][35] and are characterized by one of the highest levels of development. This results in the existence of a huge number of defined compositions with determined properties and many of them have been researched or designed specifically to function in biomedical applications. The most recent ones, including bulk materials and thin films, are characterized by good corrosion resistance in different simulated environments and/or good in vitro/in vivo biocompatibility.

2.1. Mechanical Properties

As presented in Table 1, the exemplary Zr-based metallic glasses currently being researched are characterized by high strengths in the range between 1043 and 1904 MPa, a comparatively low Young’s modulus of 70–127 GPa, and moderate to high hardness reaching 443–843 HV (4.35–9.10 GPa). However, it should be noted, that the higher values of Young’s modulus are obtained mainly for magnetron-sputtered thin films. The possibility of obtaining very high strengths and hardness is perfect both for implants and surgical devices as it ensures the required durability and allows for material mass and volume reduction. For some thin films produced by magnetron sputtering, the observed hardness was even higher than those reported in Table 1 reaching about 1030 HV (11.1 GPa) [36], which enables the possibility of current materials performance improvement by coating application.
Table 1. Recent exemplary Zr-based amorphous alloys considered for biomedical applications together with their technological details and mechanical parameters: d—obtained diameter/thickness, σmax—compressive ultimate strength, E—Young’s modulus, HV—Vickers hardness, εel—elastic strain, εpl—plastic strain.
Composition (at.%) Production Method d (mm) Structure σmax (MPa) E (GPa) Hv (HV) εelpl (%) Ref.
Zr63.5−xTixAl9Fe4.5Cu23 (x = 0, 1.5, 3, 4.5, 6) Arc melting/suction casting 3–10 Fully amorphous 1580–1690 –/0.9–4.7 [37]
Zr61Ti2Cu25Al12 Arc melting/suction casting 2–10 Fully amorphous 83 [38][39]
Zr58.6Al15.4Co18.2Cu7.8 Arc melting/suction casting 10 Fully amorphous 1950 84 –/2.0 [40]
Zr55Ti3HfxCu32−xAl10 (x = 0, 1, 2, 3, 4, 5) Arc melting/suction casting 4–8 Fully amorphous 1695–1824 73–85 2.0–2.5/0–2.6 [41]
Zr40Ti15Cu10Ni10Be25 Arc melting/suction casting 3 Mainly amorphous 796 [42]
Zr50Ti5Cu10Ni10Be25 Arc melting/suction casting 3 Mainly amorphous 741 [42]
Zr40Ti15Cu10Ni5Si5Be25 Arc melting/suction casting 3 Partially amorphous 843 [42]
Zr70Ni16Cu6Al8 Arc melting/arc tilt casting 3 1500 * 70 2.2/0 [43]
Zr65−xTixCu20Al10Fe5 (x = 0, 2, 4, 6, 8) Arc melting/suction casting 2 Fully amorphous for x = 0, 2, 4, and partially amorphous for x = 6, 8 1405–1905 –/0–8.6 [44]
Zr56Cu24Al9Ni7−xTi4Fex (x = 0, 1, 3, 5, 7) Arc melting/suction casting 2 Fully amorphous for x = 0, 1, 3, and partially amorphous for x = 5, 7 1043–1709 3.9–6.3/0–5.6 [45]
Zr60+xTi2.5Al10Fe12.5-xCu10Ag5 (x = 0, 2.5, 5) Arc melting/suction casting/casting 1–2 Fully amorphous ~1660–1740 70–78 443–460 2.0–2.0/4–12 [46]
Zr55Co30Ti15 Arc melting/melt spinning 0.04 Fully amorphous [47]
Zr62Cu22Al10Fe5Dy1 Induction melting/melt spinning 0.04 Fully amorphous 96 495 [48]
Zr37Co34Cu20Ti9 Arc melting/melt spinning Fully amorphous 81 567 [49]
Zr40Ti35Ni14Nb11 Magnetron co-sputtering 0.0006 Fully amorphous 122 ~658 [50]
Zr46Ti40Ag14 Magnetron co-sputtering 0.0003 Fully amorphous 109 ~567 [51]
Zr46Ti43Al11 Magnetron co-sputtering 0.0002 Fully amorphous 127 ~520 [51]
Zr62.5Pd37.5 Magnetron sputtering Fully amorphous [52]
* Determined in a tensile test.

2.2. Corrosion Resistance

An imperative for non-resorbable implants and surgical tools is excellent corrosion resistance in the body environment strongly connected with a limited release of toxic ions. Meeting this condition paves the way for further in vitro and in vivo research. In recent studies, the corrosion resistance of Zr-based metallic glasses in varied simulated environments was investigated by potentiodynamic polarization tests. The reported environments included:
  • Phosphate Buffered Saline (PBS) [31][37][40][44][46][47][48][49][50][51],
  • Simulated Body Fluid (SBF) [38][42][45][53],
  • Artificial Blood Plasma Solution (ABP) [47][48][49],
  • Artificial Saliva Solution (ASS) [47][48][49],
  • Hank’s Balanced Saline Solution (HBSS) [47][48][49][52][54],
  • Ringer’s solution [30],
and standard environments of 0.6 mol/dm3 NaCl, 1 mol/dm3 HCl, 1 mol/dm3 H2SO4 [41] and 0.1 mol/dm3 NaCl combined with 10g/dm3 lactic acid [43]. The identified corrosion protection mechanism is connected with oxides layers formation on the surface of the Zr-based materials [29][31][55][56][57][58]. Cumulatively, the research showed excellent corrosion resistance of Zr-based metallic glasses, typically better than for pure Ti and Ti-based alloys. It was expressed by the low value of corrosion current density. However, their passive region width often tends to be low [44] for example in comparison with crystalline Ti-based alloys [37][40][42][43][45][46], which can result in pitting corrosion. The susceptibility of Zr-based metallic glasses to pitting corrosion in chloride-containing solutions [58], such as body fluids, is one of the main concerns related to their corrosion resistance aspect. As human cell membrane potential can reach about 0.1 V, it can stimulate the electrochemical corrosion of implants with low pitting corrosion resistance [59][60][61].

2.3. In Vitro Cytocompatibility

Very recently, many Zr-based glassy alloys have also been assessed in in vitro cellular research concerning their cytotoxicity which determines the next step for the broader application of this group of materials. The research with the use of MC3T3-E1 mouse preosteoblasts cells showed an excellent cytocompatibility of particular Zr-based BMGs [46][62][38] in comparison with Ti-based alloys and even PEEK polymer. Comparably excellent results were also obtained for other BMGs in direct and indirect contact with L929 mouse fibroblasts [54][63]. Another popular type of cells used for cytotoxicity verification purposes are human osteosarcoma cells (HOS), especially from the MG-63 cell lines. Zr-based metallic glasses showed little toxicity to HOS cells [47][48][49][64].

Another kind of research is related to material-blood contact and in this field, the Zr-based glassy alloys have been successfully evaluated. The research on the Zr56Al16Co28 BMG demonstrated its very low (below 0.12%) hemolytic rates indicating its compatibility with erythrocytes [30]. Moreover, the thin films of the Zr53Cu33Al9Ta5 material proved that they do not support blood cells’ adhesion and do not cause platelet aggregation leading to thrombosis [65][66][67], which is essential for the needles or stents.

2.4. Antibacterial Properties

A desirable characteristic of materials suitable for implants and medical devices is also the ability to prevent bacterial biofilm formation leading to severe complications including the necessity to remove the implant [68][69]. The often antibiotic treatment ineffectiveness makes it advantageous to obtain the antimicrobial characteristic in the material itself or by its surface modification [70]. Particular metal ions, such as Ag, Cu, or Ni, are effective in preventing the colonization by bacteria [71][72][73][74][75]. However, they can be also responsible for cytotoxic effects on organisms’ cells. Especially effective is Ag addition because of the low required ions concentration (way below the cytotoxicity level) to obtain the antibacterial activity [76]. Therefore, in general, a proper balance should be maintained between cytotoxic and antimicrobial effects by ensuring that the concentration of released ions will not exceed the organism’s tolerance limit.
As most of the Zr-based metallic glasses contain the elements suppressing the bacteria growth, they are generally expected to have a high ability to prevent bacterial biofilm formation. Recently, Zr-based metallic glasses proved their antibacterial properties  [36][63][51][40] both by the lethal to bacteria ions release [77][78][79]. Tand smooth non-wettable surface [77][78][80], which can also be a way to prevent the biofilm formation [69]. As metallic glasses are characterized by an utterly homogenous internal structure without grain boundaries, the achievable surface roughness is exceptionally low [81][82][83][84][85]. Together, this provides an interesting proposition for coatings of surgical tools, stents, fixators, and other biomedical devices in which cells’ assimilation is not desirable.

2.5 In Vivo Biocompatibility

In vivo animal research represents the next step in the biomedical characterization of potential candidates for wider use, as in the in vitro test it is not possible to observe the full spectrum of mutual material-organism interactions. Zr-based metallic with varied compositions showed no adverse effects after implantation into the organisms of rabbits [54], mice [30], and rats [43][86][38]. Together with possible excellent osseointegration properties [43][38] and faster recovery [38], the research results show Zr-based metallic glasses as promising candidates for bone implants. 

Moreover, recent applicational studies on the MG thin films showed that particular compositions (e.g., Zr53Cu33Al9Ta5) can be very effective for coatings of medical devices such as syringe needles [66][87]. The film improved the durability, hemocompatibility, and reduced the cells’ adhesion which resulted in a less invasive procedure (lower retraction forces) and less probability of causing thrombosis.

2.6. Summary

The described properties and functionality of Zr-based metallic glasses give hope for their fast implementation in biomedical applications to reduce the percentage of unsuccessful treatments and discomfort of patients. This is supported by their usual good mechanocompatibility with bones illustrated by a low Young’s modulus. Novel Zr-based metallic glasses also showed high strengths and hardness. They are also characterized by very high elastic strains, however, sometimes without further noticeable plasticity. These materials also possess excellent corrosion resistance in terms of corrosion current density. However, they often lack pitting corrosion resistance. Described materials also exhibited cytocompatibility with various cells better or on the same level as currently used materials. It is connected with low ions release through an insulating oxides layer, and lack of toxicity of main components. However, it was shown that suboptimal composition can lead to faster corrosion and ions release lowering the cells’ viability. For some compositions, hemocompatibility and antibacterial properties were confirmed. Recent animal studies also confirmed a lack of negative response for studied Zr-based metallic glasses after implantation. Their use results in better osseointegration, faster blood vessels formation, and most important, a less painful and faster recovery. Moreover, the treatments to improve the mentioned parameters were also recently reported. However, it should also be noted that varied alloying compositions can exhibit very different behavior despite minor elemental changes. Recent research on Zr-based metallic glasses showed that their achievable mechanical and biological properties are suitable for prospective bone implants with good osseointegration and significantly reduced stress shielding effect, temporary fixtures with low bone affiliation, stents with a low probability to cause thrombosis, needles causing low tissue damage, or surgical tools with antimicrobial properties and increased lifetime.

3. Ti-Based Metallic Glasses

Ti-based materials are another group of metallic glasses that has recently gained much attention. This metal and its alloys have been used in the biomedical industry on a wide scale for many years [88][89] due to their nearly perfect cytocompatibility [90], exceptional corrosion resistance (connected with stable TiO2 oxides formation) [3], very low achievable Young’s modulus—even lower than for the bones—and excellent osseointegration [88]. Moreover, Ti does not play any biological role in the human body and no toxic effects are observed even after taking large doses of this element [1]. Nevertheless, Ti-based alloys also face problems mainly related to wear resistance and fatigue strength [1][2][3][91][92].
After the discovery of metallic glasses, Ti became one of the most widespread and beneficial elements in their compositional design, often in connection with Zr. Therefore, with its excellent biological properties, Ti has formed the basis of a new group of metallic glasses for biomedical applications. The most researched materials are the TiZrCuPd and TiZrCuFeSn compositions with their modifications.

3.1. Mechanical Properties

From the data gathered in Table 2, it is visible that Ti-based metallic glasses with various properties-modifying additions are characterized by remarkably high ultimate strengths in the range of 1261–3200 MPa, whereby the lowest values are obtained for materials produced by powder metallurgy methods [93]. These values are much higher than the compressive ultimate stress of 42 to 205 MPa for human femoral cortical bone depending on the direction of measurement. Such high values of ultimate strength are a guarantee of the durability of implants or surgical tools even in extreme situations. Simultaneously, the materials are not very brittle and generally possess distinctive plastic strain. The hardness is very uniform between varied materials falling within the range of 454–734 HV (4.9 to 7.2 GPa) with the majority of the result between 500 and 600 HV. Similar uniformity is visible for values of Young’s modulus, which ranges between 79.7 and 174 GPa with most values being about 90 to 100 GPa.
Table 2. Recent exemplary Ti-based amorphous alloys considered for biomedical applications with their technological details and mechanical parameters: d—obtained diameter/thickness, σmax—compressive ultimate strength, E—Young’s modulus, HV—Vickers hardness, εel—elastic strain, εpl—plastic strain.
Composition (at.%) Production Method d (mm) Structure σmax (MPa) E (GPa) Hv (HV) εelpl (%) Ref.
(Ti55Zr15Be20Ni10)100−xFex (x = 0, 2, 4, 6, 8, 10) Arc melting/suction casting 5–10 Fully amorphous 1878–2355 –/3.4–1.3 [94]
Ti47Cu38Zr7.5Fe2.5Sn2Si1Ag2 Arc melting/tilt pouring 7 Fully amorphous 2080 100 588 –/2.5 [95]
Ti48Cu37Zr7.5Fe2.5Sn2Si1Ag2 Arc melting/tilt pouring 6 Fully amorphous 2050 101 571 2.0/2.8 [96]
Ti45Zr10Cu31Pd10Sn4 Argon atomization/spark plasma sintering 6 Fully amorphous 100 [97]
Ti47−xCu40Zr7.5Fe2.5Sn2Si1Scx (x = 0, 1, 2, 3, 4) Arc melting/suction casting/tilt pouring 3–6 Fully amorphous 1982–2169 93–101 577–590 –/0.8–5.9 [98]
Ti47Cu38−xZr7.5Fe2.5Sn2Si1Ag2Nbx (x = 0, 1, 2) Arc melting/suction casting 3–5 Fully amorphous 2031–2078 97–100 588–593 –/1.9–2.5 [99]
Ti40Zr10Cu36Pd14 Arc melting/tilt pouring 5 Fully amorphous 2010 96 556 2.0/0.7 [100][101]
Ti40Zr10Cu36−xPd14Gax (x = 2, 4, 8, 10) Arc melting/suction casting 3 Fully amorphous for x = 2, 4, and partially amorphous for x = 8, 10 1935–2075 93–140 1.9–2.1/0.8–2.5 [102]
Ti47Cu40−xZr7.5Fe2.5Sn2Si1Tax (x = 1, 2, 3, 4) Arc melting/suction casting 3 Fully amorphous 2041–2191 98–101 582–595 –/1.0–3.4 [103]
Ti40Zr35Cu17S8 Induction melting/arc melting/suction casting 3 Fully amorphous 3200 96 509 [104]
Ti50Zr25Cu17S8 Induction melting/arc melting/suction casting 2 Fully amorphous 3100 98 524 [104]
Ti40Zr10Cu38Pd12 Induction melting/mold casting 2 Fully amorphous 2300 95 734 –/4.0 [105]
Ti40Zr10Cu34Pd14Sn2 Arc melting/suction casting 1.5 Fully amorphous >2000 93 2.2/– [106]
Ti60Zr15Cu17S8 Induction melting/arc melting/suction casting 1 Fully amorphous 2800 98 547 [104]
TiCuZrPd:Bx (x = 0, 4, 8, 14) Pulsed laser deposition Fully amorphous 108–174 454–685 [107]
Ti42Zr35Ta3Si5Co12.5Sn2.5 Argon atomization/hot pressing Fully amorphous 1261 79.7 [93]

3.2. Corrosion Resistance

As crystalline Ti-based alloys already show very good corrosion resistance, the expectations and demands from Ti-based metallic glasses are high. Recently, many different compositions were evaluated in various standard and bio-related environments including:
  • Phosphate Buffered Saline (PBS) [95][98][105][108],
  • Borate Buffered Solution with 0.1 M NaCl (BBS) [104],
  • Simulated Body Fluid (SBF) [107][109],
  • Hank’s Balanced Saline Solution (HBSS) [59][97][99][103][110],
  • 3.5 wt.% NaCl [94][111][112],
  • 0.9 wt.% NaCl [101][103][113][114].
On this basis, it can be stated that in general, Ti-based metallic glasses possess better corrosion resistance in terms of corrosion current density than commercially pure Ti and better or similar to the reference Ti-6Al-4V alloy. Similarly, to Zr-based metallic glasses the susceptibility to pitting corrosion in chloride-containing solutions is also noticeable [59][94][95][96][98][99][101][103][105][108][111][112][113]. Still, it is not as pronounced, and the passive regions are wider. In general, Ti-based metallic glasses tend to have better pitting corrosion resistance and wider passive regions than Zr-based materials. However, they are also characterized by slightly higher corrosion current densities in corresponding solutions.

3.3 In Vitro Cytocompatibility

Recent cytocompatibility research for the Ti-based metallic glasses was mainly conducted with the use of the MC3T3-E1 mouse preosteoblasts cell line. Varied materials outperformed even the reference Ti-6Al-4V material [99][98]. or reached a similar level of cytocompatibility [95][96]. Similarly, good results were obtained in tests of Ti-based metallic glasses with the use of Saos-2 human osteosarcoma cells [115][116][109], human gingival fibroblasts [100], MG-63 human osteosarcoma cells [101], and L929 mouse fibroblasts [117][109][54]. Materials with confirmed in vitro cytocompatibility are good candidates for further biomedical applications.

In terms of interactions with blood, the Ti-based metallic glasses also showed superior hemocompatibility. The TiNbZrSi alloy coating was found to be non-hemolytic and, unlike the uncoated Ti-6Al-4V sample, does not exhibit platelet adherence and stimulation leading to thrombus formation [109]. The same results were also obtained for the thin films based on TiCuZrPd alloy with B addition [107].

3.4. Antibacterial Properties

There is little recent research on the antimicrobial properties of Ti-based metallic glasses. However, the available ones showed that the porous Ti45Zr10Cu31Pd10Sn4 BMG is effective in the long term in restricting the growth of S. aureus [118]. It was confirmed that induced porosity leads to a higher concentration of Cu ions, providing the antibacterial effect, while still being below the cytotoxicity threshold of MC3T3-E1 cells.
Other research on Ti40Zr10Cu36Pd14 BMG confirmed its antibacterial activity against Aggregatibacter actinomycetemcomitans, which is common in the oral environment and responsible for periodontal and peri-implant diseases [100]. Moreover, a significant reduction in multispecies biofilm formation was observed for the examined BMG in comparison to the Ti-6Al-4V sample, which was fully covered in a thick layer of different microorganisms present in human oral flora. The antimicrobial action of BMG was connected with a non-wettable surface preventing bacteria adherence and the presence of Cu in the surface layer, similarly as for Zr-based materials, confirming the Ti40Zr10Cu36Pd14 BMG as a viable candidate for dental implants.
Based on this and the results for the Zr-based materials, the antimicrobial activity is predicted for other recently researched Ti-based metallic glasses with the content of elements inhibiting bacteria growth such as Cu or Ag, with a non-wettable surface being an additional advantage, however, more research is needed in this direction.

3.5. In Vivo Biocompatibility

In the field of in vivo animal research performed with Ti-based metallic glasses, there are also successful applications as implants. No inflammatory reactions nor necrosis or toxicity symptoms were observed after bone implantations in rats [106][59][97] and rabbits [119] indicating the preliminary signs of safety in short-term biomedical applications. Particular Ti-based chemical compositions of metallic glasses showed also excellent osseointegration properties [106][97] and induction of new bone tissue growth [119]. These properties are especially important when considering an application for load-bearing bone implants.

3.6. Summary

Based on available research, it can be stated that Ti-based metallic glasses possess very favorable properties for bio-related devices. However, their overall development level is slightly below the one for Zr-based metallic glasses. Due to their mechanical properties, corrosion resistance, hemo- and biocompatibility, they may find application as: bulk materials or coatings for implants with particularly good osseointegration and better than current Ti alloys wear resistance, surgical tools with very high strength, or stents without the proneness to cause platelets aggregation. Not without significance is the proven possibility of controlling their mechanical and structural properties by powder metallurgy to reach the bones’ Young’s modulus and cells’ ingrowth capabilities. Their superior to Zr-based metallic glasses pitting corrosion resistance and often better cytocompatibility is a good sign for prolonged use. As Ti-based metallic glasses also often contain fewer potentially harmful elements, they are good candidates for permanent applications. In spite of the single excellent results, the area of antibacterial properties research and in vivo testing is less developed than in the case of Zr-based amorphous alloys, which have already proven their exceptional performance up to a certain point. However, also in the case of Ti-based metallic glasses, some materials were assessed with no negative in vivo response.

4. Other Metallic Glasses

4.1. Mg-Based Metallic Glasses

Contrary to all described above materials which are meant to be as corrosion resistant as possible to prevent the ions’ release, toxicity, and device damage, the Mg-based metallic glasses are driven by a different philosophy. According to the third generation of biomaterials guidelines which include the ability of a material to trigger specific cellular response [3], the implanted materials are intended to be temporary structures that do not require further surgical removal, and that support the regeneration processes of the organism’s tissues [1]. For this, they are designed to be degradable to allow the native tissue integration and gradual replacement of implant [1][120]. Such an approach enforces the use of only completely bio-friendly components, such as Mg, which is the essential ingredient in the human body and can be absorbed and removed by the organism without causing adverse effects. Recently, the very dynamic development of this material group can be observed with many very similar compositions [121].

4.2. Ta-Based Metallic Glasses

Tantalum is one of the highly desirable metals to use in biomedical applications due to its excellent biocompatibility and corrosion resistance connected with good mechanical properties [1]. It is used as a minor alloying element in Zr- or Ti-based metallic glasses. However, Ta-based metallic glasses are difficult to manufacture due to the poor GFA. There are very few results described in the literature. The reported Ta42Ni40Co18 BMG with a 2 mm critical diameter also contains potentially harmful Ni and Co and possesses remarkably high strength (2.7 GPa), and a relatively high Young’s modulus (170 GPa) [122]. This directs its possible application to surgical tools or stents. However, it was reported that the creation of Ta-based TaZrCuAlAg thin film metallic glasses by magnetron co-sputtering is possible [123]. Their hardness is about 7.5 GPa (~695 HV), and the Young’s modulus is 130 to 145 GPa depending on composition indicating a possible application for coatings. Thin films metallic glass compositions of Ta57Ti17Zr15Si11 and Ta75Ti10Zr8Si7, without potentially harmful elements, already demonstrated their excellent cytocompatibility with D1 mouse mesenchymal cells [124].

4.3. Pd-Based Metallic Glasses

Due to the biocompatibility of Pd [125][126] and the remarkably high GFA of compositions in which it is present [17] recent biomedical research also includes materials based on this element. It was shown that Pd40Cu30Ni10P20 BMG [127] exhibits a hardness of about 500 HV which is more than for popular biomedical materials such as 316L steel, Ti-6Al-4V alloy, and even CoCrMo alloy. Moreover, it possesses exceptional wear resistance both in dry conditions and corrosive phosphate buffer saline (PBS) which is about 30 to 40 times higher than for Ti-6Al-4V alloy. Its corrosion resistance measured by corrosion current density is also superior to Ti-6Al-4V, indicating lower corrosion rates. All these characteristics are significant for biomedical applications. Yet the disadvantage is the presence of potentially toxic Cu and Ni in the described BMG composition, which can be problematic in further in vitro studies.
The Pd77.5Si16.5Cu6 composition [128] with a smaller number of harmful elements, very recently demonstrated excellent hemocompatibility and thrombogenic resistance preventing the platelets aggregation and activation in comparison to Ti-6Al-4V alloy. This results in the possible application of this BMG for stents or blood-pumping devices with a low risk of causing thrombosis.

References

  1. Chen, Q.; Thouas, G.A. Metallic Implant Biomaterials. Mater. Sci. Eng. R Rep. 2015, 87, 1–57.
  2. Ibrahim, M.Z.; Sarhan, A.A.D.; Yusuf, F.; Hamdi, M. Biomedical Materials and Techniques to Improve the Tribological, Mechanical and Biomedical Properties of Orthopedic Implants—A Review Article. J. Alloy. Compd. 2017, 714, 636–667.
  3. Navarro, M.; Michiardi, A.; Castaño, O.; Planell, J.A. Biomaterials in Orthopaedics. J. R. Soc. Interface 2008, 5, 1137–1158.
  4. Huiskes, R.; 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.
  5. Morgan, E.F.; Unnikrisnan, G.U.; Hussein, A.I. Bone Mechanical Properties in Healthy and Diseased States. Annu. Rev. Biomed. Eng. 2018, 20, 119–143.
  6. Niinomi, M.; Nakai, M. Titanium-Based Biomaterials for Preventing Stress Shielding between Implant Devices and Bone. Int. J. Biomater. 2011, 2011, 836587.
  7. Capizzani, R. Magnetic Resonance Imaging Hazards and Safety Guidelines. Strateg. Outcomes Pract.-Tech. Advis. Bull. 2009, 1–11.
  8. Chockattu, S.J.; Suryakant, D.B.; Thakur, S. Unwanted Effects Due to Interactions between Dental Materials and Magnetic Resonance Imaging: A Review of the Literature. Restor. Dent. Endod. 2018, 43, e39.
  9. Buckwalter, K.; Lin, C.; Ford, J. Managing Postoperative Artifacts on Computed Tomography and Magnetic Resonance Imaging. Semin. Musculoskelet. Radiol. 2011, 15, 309–319.
  10. Zhao, B.; Yang, B.; Rodríguez-Viejo, J.; Wu, M.; Schick, C.; Zhai, Q.; Gao, Y. Bridging the Local Configurations and Crystalline Counterparts of Bulk Metallic Glass by Nanocalorimetry. J. Mater. Res. Technol. 2019, 8, 3603–3611.
  11. Kramer, J. The Amorphous State of Metals. Z. Für Phys. 1936, 106, 675–691.
  12. Cheng, Y.-T.; Johnson, W.L. Disordered Materials: A Survey of Amorphous Solids. Science 1987, 235, 997–1002.
  13. Klement, W.; Willens, R.H.; Duwez, P. Non-Crystalline Structure in Solidified Gold–Silicon Alloys. Nature 1960, 187, 869–870.
  14. Schroers, J. Bulk Metallic Glasses. Phys. Today 2013, 66, 32–37.
  15. Chen, H.S. Thermodynamic Considerations on the Formation and Stability of Metallic Glasses. Acta Metall. 1974, 22, 1505–1511.
  16. Gu, J.-L.; Shao, Y.; Yao, K.-F. The Novel Ti-Based Metallic Glass with Excellent Glass Forming Ability and an Elastic Constant Dependent Glass Forming Criterion. Materialia 2019, 8, 100433.
  17. Nishiyama, N.; Takenaka, K.; Miura, H.; Saidoh, N.; Zeng, Y.; Inoue, A. The World’s Biggest Glassy Alloy Ever Made. Intermetallics 2012, 30, 19–24.
  18. Fornell, J.; van Steenberge, N.; Varea, A.; Rossinyol, E.; Pellicer, E.; Suriñach, S.; Baró, M.D.; Sort, J. Enhanced Mechanical Properties and in Vitro Corrosion Behavior of Amorphous and Devitrified Ti40Zr10Cu38Pd12 Metallic Glass. J. Mech. Behav. Biomed. Mater. 2011, 4, 1709–1717.
  19. Hasiak, M.; Tkaczyk, M.; Łaszcz, A.; Olszewski, J. Effect of Alloying Additions on Microstructure, Mechanical and Magnetic Properties of Rapidly Cooled Bulk Fe-B-M-Cu (M = Ti, Mo and Mn) Alloys. Metall. Mater. Trans. A Phys. Metall. Mater. Sci. 2022, 53, 556–572.
  20. Hasiak, M.; Świerczek, J. Some Thermomagnetic and Mechanical Properties of Amorphous Fe75Zr4Ti3Cu1B17 Ribbons. Materials 2022, 15, 368.
  21. 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.
  22. Świerczek, J.; Kupczyk, A.; Hasiak, M. Impact of Annealing on Flattening of Magnetic Entropy Change versus Temperature Curves in Amorphous and Partially Crystallized Fe76Mo10Cu1B13 Alloy. Acta Phys. Pol. A 2019, 135, 212–214.
  23. Xie, C.; Milošev, I.; Renner, F.U.; Kokalj, A.; Bruna, P.; Crespo, D. Corrosion Resistance of Crystalline and Amorphous CuZr Alloys in NaCl Aqueous Environment and Effect of Corrosion Inhibitors. J. Alloy. Compd. 2021, 879, 160464.
  24. Telford, M. The Case for Bulk Metallic Glass. Mater. Today 2004, 7, 36–43.
  25. Greer, A.L. Metallic Glasses…on the Threshold. Mater. Today 2009, 12, 14–22.
  26. Ashby, M.; Greer, A. Metallic Glasses as Structural Materials. Scr. Mater. 2006, 54, 321–326.
  27. Suryanarayana, C.; Inoue, A. Bulk Metallic Glasses, 2nd ed.; CRC Press: Boca Raton, FL, USA, 2017; ISBN 9781315153483.
  28. Green, B.A.; Liaw, P.; Buchanan, R.A. Corrosion Behavior. In Bulk Metallic Glasses; Miller, M., Liaw, P., Eds.; Springer: New York, NY, USA, 2008; pp. 205–234.
  29. Scully, J.R.; Gebert, A.; Payer, J.H. Corrosion and Related Mechanical Properties of Bulk Metallic Glasses. J. Mater. Res. 2007, 22, 302–313.
  30. Wang, R.; Wang, Y.; Yang, J.; Sun, J.; Xiong, L. Influence of Heat Treatment on the Mechanical Properties, Corrosion Behavior, and Biocompatibility of Zr56Al16Co28 Bulk Metallic Glass. J. Non Cryst. Solids 2015, 411, 45–52.
  31. 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.
  32. Gan, Z.; Zhang, C.; Zhang, Z.-R.; Chen, Z.-J.; Liu, L. Crystallization-Dependent Transition of Corrosion Resistance of an Fe-Based Bulk Metallic Glass under Hydrostatic Pressures. Corros. Sci. 2021, 179, 109098.
  33. Ikarashi, Y.; Toyoda, K.; Kobayashi, E.; Doi, H.; Yoneyama, T.; Hamanaka, H.; Tsuchiya, T. Improved Biocompatibility of Titanium-Zirconium (Ti-Zr) Alloy: Tissue Reaction and Sensitization to Ti-Zr Alloy Compared with Pure Ti and Zr in Rat Implantation Study. J. Jpn. Inst. Met. 2007, 71, 395–401.
  34. Peker, A.; Johnson, W.L. A Highly Processable Metallic Glass: Zr 41.2 Ti 13.8 Cu 12.5 Ni 10.0 Be 22.5. Appl. Phys. Lett. 1993, 63, 2342–2344.
  35. Qiao, D.; Peker, A. Enhanced Glass Forming Ability in Zr-Based Bulk Metallic Glasses with Hf Addition. Intermetallics 2012, 24, 115–119.
  36. 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.
  37. Han, K.-M.; Jiang, H.; Wang, Y.-M.; Qiang, J.-B. Zr–Ti–Al–Fe–Cu Bulk Metallic Glasses for Biomedical Device Application. Rare Met. 2021, 40, 1239–1246.
  38. Sun, K.; Fu, R.; Liu, X.W.; Xu, L.M.; Wang, G.; Chen, S.Y.; Zhai, Q.J.; Pauly, S. Osteogenesis and Angiogenesis of a Bulk Metallic Glass for Biomedical Implants. Bioact. Mater. 2022, 8, 253–266.
  39. He, Q.; Cheng, Y.Q.; Ma, E.; Xu, J. Locating Bulk Metallic Glasses with High Fracture Toughness: Chemical Effects and Composition Optimization. Acta Mater. 2011, 59, 202–215.
  40. Han, K.; Jiang, H.; Wang, Y.; Qiang, J.; Yu, C. Antimicrobial Zr-Based Bulk Metallic Glasses for Surgical Devices Applications. J. Non Cryst. Solids 2021, 564, 120827.
  41. Jin, Z.S.; Yang, Y.J.; Zhang, Z.P.; Ma, X.Z.; Lv, J.W.; Wang, F.L.; Ma, M.Z.; Zhang, X.Y.; Liu, R.P. Effect of Hf Substitution Cu on Glass-Forming Ability, Mechanical Properties and Corrosion Resistance of Ni-Free Zr–Ti–Cu–Al Bulk Metallic Glasses. J. Alloy. Compd. 2019, 806, 668–675.
  42. Hasiak, M.; Sobieszczańska, B.; Łaszcz, A.; Biały, M.; Chęcmanowski, J.; Zatoński, T.; Bożemska, E.; Wawrzyńska, M. Production, Mechanical Properties and Biomedical Characterization of ZrTi-Based Bulk Metallic Glasses in Comparison with 316L Stainless Steel and Ti6Al4V Alloy. Materials 2021, 15, 252.
  43. Ida, H.; Seiryu, M.; Takeshita, N.; Iwasaki, M.; Yokoyama, Y.; Tsutsumi, Y.; Ikeda, E.; Sasaki, S.; Miyashita, S.; Sasaki, S.; et al. Biosafety, Stability, and Osteogenic Activity of Novel Implants Made of Zr70Ni16Cu6Al8 Bulk Metallic Glass for Biomedical Application. Acta Biomater. 2018, 74, 505–517.
  44. Shi, H.; Zhao, W.; Wei, X.; Ding, Y.; Shen, X.; Liu, W. Effect of Ti Addition on Mechanical Properties and Corrosion Resistance of Ni-Free Zr-Based Bulk Metallic Glasses for Potential Biomedical Applications. J. Alloy. Compd. 2020, 815, 152636.
  45. Shi, H.; Zhou, H.; Zhou, Z.; Ding, Y.; Liu, W.; Ji, J. Improved Mechanical Properties and Corrosion Resistance of Zr-Cu-Al-Ni-Ti Bulk Metallic Glass by Fe Substitution for Ni. J. Non Cryst. Solids 2022, 576, 121246.
  46. 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.
  47. Vincent, S.; Daiwile, A.; Devi, S.S.; Kramer, M.J.; Besser, M.F.; Murty, B.S.; Bhatt, J. Bio-Corrosion and Cytotoxicity Studies on Novel Zr55Co30Ti15 and Cu60Zr20Ti20 Metallic Glasses. Metall. Mater. Trans. A 2015, 46, 2422–2430.
  48. Jain, A.; Prabhu, Y.; Gunderov, D.; Narayan, R.L.; Saini, P.; Vincent, S.; Sudha, P.; Bagde, A.D.; Bhatt, J. Structural Characterization, Biocorrosion and in-Vitro Investigation on Zr62Cu22Al10Fe5Dy1 Metallic Glass for Bio-Implant Applications. J. Non Cryst. Solids 2022, 598, 121928.
  49. Prabhu, Y.; Jain, A.; Vincent, S.; Ryu, W.H.; Park, E.S.; Kumar, R.; Bagde, A.D.; Bhatt, J. Compositional Design and in Vitro Investigation on Novel Zr–Co–Cu–Ti Metallic Glass for Biomedical Applications. Intermetallics 2022, 150, 107692.
  50. 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.
  51. 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.
  52. 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.
  53. Rajan, S.T.; V, A.T.V.; Terada-Nakaishi, M.; Chen, P.; Hanawa, T.; Nandakumar, A.K.; Subramanian, B. Zirconium-Based Metallic Glass and Zirconia Coatings to Inhibit Bone Formation on Titanium. Biomed. Mater. 2020, 15, 065019.
  54. 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.
  55. Liu, L.; Qiu, C.L.; Chen, Q.; Chan, K.C.; Zhang, S.M. Deformation Behavior, Corrosion Resistance, and Cytotoxicity of Ni-Free Zr-Based Bulk Metallic Glasses. J. Biomed. Mater. Res. Part A 2008, 86, 160–169.
  56. Gebert, A.; Buchholz, K.; Leonhard, A.; Mummert, K.; Eckert, J.; Schultz, L. Investigations on the Electrochemical Behaviour of Zr-Based Bulk Metallic Glasses. Mater. Sci. Eng. A 1999, 267, 294–300.
  57. 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.
  58. Wang, B.; Xu, K.K.; Shi, X.H.; Zhang, M.; Qiao, J.W.; Gao, C.H.; Wu, Y.C. Electrochemical and Chemical Corrosion Behaviors of the In-Situ Zr-Based Metallic Glass Matrix Composites in Chloride-Containing Solutions. J. Alloy. Compd. 2019, 770, 679–685.
  59. Lin, C.H.; Chen, C.H.; Huang, Y.S.; Huang, C.H.; Huang, J.C.; Jang, J.S.C.; Lin, Y.S. In-Vivo Investigations and Cytotoxicity Tests on Ti/Zr-Based Metallic Glasses with Various Cu Contents. Mater. Sci. Eng. C 2017, 77, 308–317.
  60. Kadir, L.A.; Stacey, M.; Barrett-Jolley, R. Emerging Roles of the Membrane Potential: Action beyond the Action Potential. Front. Physiol. 2018, 9, 1661.
  61. Wu, J.; Li, M.; Zhou, Y.; Xu, K.; Li, X.; Lin, C.; Zhang, J.; Cai, K. Pitting Behavior of 316L Stainless Steel in Direct Culture with Mesenchymal Stem Cells. Corros. Sci. 2022, 204, 110380.
  62. Du, C.; Yang, Y.; Zheng, L.; Zhang, T.; Zhao, X.; Wang, C. Structure-Element Surface Modification Strategy Enhances the Antibacterial Performance of Zr-BMGs. ACS Appl. Mater. Interfaces 2022, 14, 8793–8803.
  63. Subramanian, B.; Maruthamuthu, S.; Rajan, S.T. Biocompatibility Evaluation of Sputtered Zirconium-Based Thin Film Metallic Glass-Coated Steels. Int. J. Nanomed. 2015, 10, 17–29.
  64. Pang, H.; Xu, C.; Qin, H.; Li, D.; Li, P.; Wang, B.; Zhang, S.; Zhao, Q. Cytotoxicity in Vitro of a Noval Ni-Free ZrCuFeAlAg Bulk Metallic Glass. J. Biomed. Eng. 2015, 32, 380–386.
  65. Chang, C.H.; Li, C.L.; Yu, C.C.; Chen, Y.L.; Chyntara, S.; Chu, J.P.; Chen, M.J.; Chang, S.H. Beneficial Effects of Thin Film Metallic Glass Coating in Reducing Adhesion of Platelet and Cancer Cells: Clinical Testing. Surf. Coat. Technol. 2018, 344, 312–321.
  66. Bai, M.-Y.; Chang, Y.-C.; Chu, J.P. Preclinical Studies of Non-Stick Thin Film Metallic Glass-Coated Syringe Needles. Sci. Rep. 2020, 10, 20313.
  67. Diyatmika, W.; Yu, C.C.; Tanatsugu, Y.; Yasuzawa, M.; Chu, J.P. Fibrinogen and Albumin Adsorption Profiles on Ni-Free Zr-Based Thin Film Metallic Glass. Thin Solid Films 2019, 688, 137382.
  68. Ronin, D.; Boyer, J.; Alban, N.; Natoli, R.M.; Johnson, A.; Kjellerup, B.V. Current and Novel Diagnostics for Orthopedic Implant Biofilm Infections: A Review. APMIS 2022, 130, 59–81.
  69. Rodríguez-Merchán, E.C.; Davidson, D.J.; Liddle, A.D. Recent Strategies to Combat Infections from Biofilm-Forming Bacteria on Orthopaedic Implants. Int. J. Mol. Sci. 2021, 22, 10243.
  70. Blank, E.; Grischke, J.; Winkel, A.; Eberhard, J.; Kommerein, N.; Doll, K.; Yang, I.; Stiesch, M. Evaluation of Biofilm Colonization on Multi-Part Dental Implants in a Rat Model. BMC Oral. Health 2021, 21, 313.
  71. Chen, Y.; Ding, Y.; Zheng, J. A Polymer Nanocomposite Coating with Enhanced Hydrophilicity, Antibacterial and Antibiofouling Properties: Role of Polymerizable Emulsifier/Anionic Ligand. Chem. Eng. J. 2020, 379, 122268.
  72. Wang, X.; Li, R.; Liu, A.; Yue, C.; Wang, S.; Cheng, J.; Li, J.; Liu, Z. Syntheses, Crystal Structures, Antibacterial Activities of Cu(II) and Ni(II) Complexes Based on Terpyridine Polycarboxylic Acid Ligand. J. Mol. Struct. 2019, 1184, 503–511.
  73. Vimbela, G.V.; Ngo, S.M.; Fraze, C.; Yang, L.; Stout, D.A. Antibacterial Properties and Toxicity from Metallic Nanomaterials. Int. J. Nanomed. 2017, 12, 3941–3965.
  74. Liu, W.; Li, J.; Cheng, M.; Wang, Q.; Qian, Y.; Yeung, K.W.K.; Chu, P.K.; Zhang, X. A Surface-Engineered Polyetheretherketone Biomaterial Implant with Direct and Immunoregulatory Antibacterial Activity against Methicillin-Resistant Staphylococcus Aureus. Biomaterials 2019, 208, 8–20.
  75. Grass, G.; Rensing, C.; Solioz, M. Metallic Copper as an Antimicrobial Surface. Appl. Environ. Microbiol. 2011, 77, 1541–1547.
  76. Melaiye, A.; Youngs, W.J. Silver and Its Application as an Antimicrobial Agent. Expert Opin. Ther. Pat. 2005, 15, 125–130.
  77. Tang, J.F.; Huang, P.Y.; Lin, J.H.; Liu, T.W.; Yang, F.C.; Chang, C.L. Microstructure and Antimicrobial Properties of Zr-Cu-Ti Thin-Film Metallic Glass Deposited Using High-Power Impulse Magnetron Sputtering. Materials 2022, 15, 2461.
  78. Etiemble, A.; der Loughian, C.; 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. Alloy. Compd. 2017, 707, 155–161.
  79. Huang, L.; Fozo, E.M.; Zhang, T.; Liaw, P.K.; He, W. Antimicrobial Behavior of Cu-Bearing Zr-Based Bulk Metallic Glasses. Mater. Sci. Eng. C 2014, 39, 325–329.
  80. Sharifikolouei, E.; Najmi, Z.; Cochis, A.; Scalia, A.C.; Aliabadi, M.; Perero, S.; Rimondini, L. Generation of Cytocompatible Superhydrophobic Zr–Cu–Ag Metallic Glass Coatings with Antifouling Properties for Medical Textiles. Mater. Today Bio. 2021, 12, 100148.
  81. Zhao, J.; Gao, M.; Ma, M.; Cao, X.; He, Y.; Wang, W.; Luo, J. Influence of Annealing on the Tribological Properties of Zr-Based Bulk Metallic Glass. J. Non Cryst. Solids 2018, 481, 94–97.
  82. Loye, A.M.; Kinser, E.R.; Bensouda, S.; Shayan, M.; Davis, R.; Wang, R.; Chen, Z.; Schwarz, U.D.; Schroers, J.; Kyriakides, T.R. Regulation of Mesenchymal Stem Cell Differentiation by Nanopatterning of Bulk Metallic Glass. Sci. Rep. 2018, 8, 8758.
  83. Liu, L.; Zhang, T.; Liu, Z.; Yu, C.; Dong, X.; He, L.; Gao, K.; Zhu, X.; Li, W.; Wang, C.; et al. Near-Net Forming Complex Shaped Zr-Based Bulk Metallic Glasses by High Pressure Die Casting. Materials 2018, 11, 2338.
  84. Rauf, A.; Guo, C.Y.; Fang, Y.N.; Yu, Z.; Sun, B.A.; Feng, T. Binary Cu-Zr Thin Film Metallic Glasses with Tunable Nanoscale Structures and Properties. J. Non Cryst. Solids 2018, 498, 95–102.
  85. Jiao, Y.; Brousseau, E.; Shen, X.; Wang, X.; Han, Q.; Zhu, H.; Bigot, S.; He, W. Investigations in the Fabrication of Surface Patterns for Wettability Modification on a Zr-Based Bulk Metallic Glass by Nanosecond Laser Surface Texturing. J. Mater. Process. Technol. 2020, 283, 116714.
  86. Imai, K.; Hiromoto, S. In Vivo Evaluation of Bulk Metallic Glasses for Osteosynthesis Devices. Materials 2016, 9, 676.
  87. Chu, J.P.; Yu, C.C.; Tanatsugu, Y.; Yasuzawa, M.; Shen, Y.L. Non-Stick Syringe Needles: Beneficial Effects of Thin Film Metallic Glass Coating. Sci. Rep. 2016, 6, 31847.
  88. Khorasani, A.M.; Goldberg, M.; Doeven, E.H.; Littlefair, G. Titanium in Biomedical Applications—Properties and Fabrication: A Review. J. Biomater. Tissue Eng. 2015, 5, 593–619.
  89. Zhang, L.C.; Chen, L.Y. A Review on Biomedical Titanium Alloys: Recent Progress and Prospect. Adv. Eng. Mater. 2019, 21, 1801215.
  90. Park, Y.-J.; Song, Y.-H.; An, J.-H.; Song, H.-J.; Anusavice, K.J. Cytocompatibility of Pure Metals and Experimental Binary Titanium Alloys for Implant Materials. J. Dent. 2013, 41, 1251–1258.
  91. Hua, N.; Wang, W.; Wang, Q.; Ye, Y.; Lin, S.; Zhang, L.; Guo, Q.; Brechtl, J.; Liaw, P.K. Mechanical, Corrosion, and Wear Properties of Biomedical Ti–Zr–Nb–Ta–Mo High Entropy Alloys. J. Alloy. Compd. 2021, 861, 157997.
  92. Hussein, M.A.; Mohammed, A.S.; Al-Aqeeli, N. Wear Characteristics of Metallic Biomaterials: A Review. Materials 2015, 8, 2749–2768.
  93. Wong, P.C.; Song, S.M.; Tsai, P.H.; Maqnun, M.J.; Wang, W.R.; Wu, J.L.; Jang, S.-C.C. Using Cu as a Spacer to Fabricate and Control the Porosity of Titanium Zirconium Based Bulk Metallic Glass Foams for Orthopedic Implant Applications. Materials 2022, 15, 1887.
  94. Gu, J.; Yang, X.; Zhang, A.; Shao, Y.; Zhao, S.; Yao, K. Centimeter-Sized Ti-Rich Bulk Metallic Glasses with Superior Specific Strength and Corrosion Resistance. J. Non Cryst. Solids 2019, 512, 206–210.
  95. 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. Alloy. Compd. 2015, 625, 323–327.
  96. Yang, W.; Liu, Y.; Hua, N.; Pang, S.; Li, Y.; Liaw, P.K.; Zhang, T. Formation and Properties of Biocompatible Ti-Based Bulk Metallic Glasses in the Ti–Cu–Zr–Fe–Sn–Si–Ag System. J. Non Cryst. Solids 2021, 571, 121060.
  97. Xie, G.; Kanetaka, H.; Kato, H.; Qin, F.; Wang, W. Porous Ti-Based Bulk Metallic Glass with Excellent Mechanical Properties and Good Biocompatibility. Intermetallics 2019, 105, 153–162.
  98. Liu, Y.; Wang, G.; Li, H.; Pang, S.; Chen, K.; Zhang, T. Ti-Cu-Zr-Fe-Sn-Si-Sc Bulk Metallic Glasses with Good Mechanical Properties for Biomedical Applications. J. Alloy. Compd. 2016, 679, 341–349.
  99. Yan, H.-M.; Liu, Y.; Pang, S.-J.; Zhang, T. Glass Formation and Properties of Ti-Based Bulk Metallic Glasses as Potential Biomaterials with Nb Additions. Rare Met. 2018, 37, 831–837.
  100. Rezvan, A.; Sharifikolouei, E.; Lassnig, A.; Soprunyuk, V.; Gammer, C.; Spieckermann, F.; Schranz, W.; Najmi, Z.; Cochis, A.; Scalia, A.C.; et al. Antibacterial Activity, Cytocompatibility, and Thermomechanical Stability of Ti40Zr10Cu36Pd14 Bulk Metallic Glass. Mater. Today Bio 2022, 16, 100378.
  101. Liens, A.; Etiemble, A.; Rivory, P.; Balvay, S.; Pelletier, J.M.; Cardinal, S.; Fabrègue, D.; Kato, H.; Steyer, P.; Munhoz, T.; et al. On the Potential of Bulk Metallic Glasses for Dental Implantology: Case Study on Ti40Zr10Cu36Pd14. Materials 2018, 11, 249.
  102. Bera, S.; Ramasamy, P.; Şopu, D.; Sarac, B.; Zálešák, J.; Gammer, C.; Stoica, M.; Calin, M.; Eckert, J. Tuning the Glass Forming Ability and Mechanical Properties of Ti-Based Bulk Metallic Glasses by Ga Additions. J. Alloy. Compd. 2019, 793, 552–563.
  103. Liu, Y.; Wang, H.J.; Pang, S.J.; Zhang, T. Ti–Zr–Cu–Fe–Sn–Si–Ag–Ta Bulk Metallic Glasses with Good Corrosion Resistance as Potential Biomaterials. Rare Met. 2020, 39, 688–694.
  104. Kuball, A.; Gross, O.; Bochtler, B.; Adam, B.; Ruschel, L.; Zamanzade, M.; Busch, R. Development and Characterization of Titanium-Based Bulk Metallic Glasses. J. Alloy. Compd. 2019, 790, 337–346.
  105. Hua, N.; Hong, X.; Lin, L.; Liao, Z.; Zhang, L.; Ye, X.; Wang, Q. Mechanical, Corrosion, and Wear Performances of a Biocompatible Ti-Based Glassy Alloy. J. Non Cryst. Solids 2020, 543, 120116.
  106. Kokubun, R.; Wang, W.; Zhu, S.; Xie, G.; Ichinose, S.; Itoh, S.; Takakuda, K. In Vivo Evaluation of a Ti-Based Bulk Metallic Glass Alloy Bar. Bio-Med. Mater. Eng. 2015, 26, 9–17.
  107. Rajan, S.T.; Das, M.; Kumar, P.S.; Arockiarajan, A.; Subramanian, B. Biological Performance of Metal Metalloid (TiCuZrPd:B) TFMG Fabricated by Pulsed Laser Deposition. Colloids Surf. B Biointerfaces 2021, 202, 111684.
  108. Lin, B.; Yang, K.; Bao, X.; Liu, J.; Guo, Q.; Zhang, L.; Wang, Q.; Hua, N. Enhanced Wear, Corrosion, and Corrosive-Wear Resistance of the Biocompatible Ti-Based Bulk Metallic Glass by Oxidation Treatment. J. Non Cryst. Solids 2022, 576, 121231.
  109. Thanka Rajan, S.; Bendavid, A.; Subramanian, B. Cytocompatibility Assessment of Ti-Nb-Zr-Si Thin Film Metallic Glasses with Enhanced Osteoblast Differentiation for Biomedical Applications. Colloids Surf. B Biointerfaces 2019, 173, 109–120.
  110. Guo, Y.; Bataev, I.; Georgarakis, K.; Jorge, A.M.; Nogueira, R.P.; Pons, M.; Yavari, A.R. Ni- and Cu-Free Ti-Based Metallic Glasses with Potential Biomedical Application. Intermetallics 2015, 63, 86–96.
  111. Gu, J.; Shao, Y.; Shi, L.; Si, J.; Yao, K. Novel Corrosion Behaviours of the Annealing and Cryogenic Thermal Cycling Treated Ti-Based Metallic Glasses. Intermetallics 2019, 110, 106467.
  112. Gong, P.; Wang, D.; Zhang, C.; Wang, Y.; Jamili-Shirvan, Z.; Yao, K.; Wang, X. Corrosion Behavior of TiZrHfBeCu(Ni) High-Entropy Bulk Metallic Glasses in 3.5 Wt. % NaCl. Npj Mater. Degrad. 2022, 6, 77.
  113. Wei, Q.; Gostin, P.F.; Addison, O.; Reed, D.; Calin, M.; Bera, S.; Ramasamy, P.; Davenport, A. The Influence of Partial Replacement of Cu with Ga on the Corrosion Behavior of Ti 40 Zr 10 Cu 36 Pd 14 Metallic Glasses. J. Electrochem. Soc. 2019, 166, C485–C491.
  114. Yüce, E.; Zarazúa-Villalobos, L.; Ter-Ovanessian, B.; Sharifikolouei, E.; Najmi, Z.; Spieckermann, F.; Eckert, J.; Sarac, B. New-Generation Biocompatible Ti-Based Metallic Glass Ribbons for Flexible Implants. Mater. Des. 2022, 223, 111139.
  115. Blanquer, A.; Hynowska, A.; Nogués, C.; Ibáñez, E.; Sort, J.; Baró, M.D.; Özkale, B.; Pané, S.; Pellicer, E.; Barrios, L. Effect of Surface Modifications of Ti40Zr10Cu38Pd12 Bulk Metallic Glass and Ti-6Al-4V Alloy on Human Osteoblasts in Vitro Biocompatibility. PLoS ONE 2016, 11, e0156644.
  116. Aliyu, A.A.A.; Udomlertpreecha, S.; Medhisuwakul, M.; Panwisawas, C.; Reed, R.; Puncreobutr, C.; Khamwannah, J.; Kuimalee, S.; Yipyintum, C.; Lohwongwatana, B. A New Toxic-Free Ti40Zr10Co36Pd14 Metallic Glass with Good Biocompatibility and Surface Behaviour Comparable to Ti-6Al-4V. Mater. Des. 2022, 218, 110691.
  117. Li, T.H.; Wong, P.C.; Chang, S.F.; Tsai, P.H.; Jang, J.S.C.; Huang, J.C. Biocompatibility Study on Ni-Free Ti-Based and Zr-Based Bulk Metallic Glasses. Mater. Sci. Eng. C 2017, 75, 1–6.
  118. Du, P.; Xiang, T.; Yang, X.; Xie, G. Enhanced Mechanical and Antibacterial Properties of Cu-Bearing Ti-Based Bulk Metallic Glass by Controlling Porous Structure. J. Alloy. Compd. 2022, 904, 164005.
  119. Liao, Y.C.; Song, S.M.; Li, T.H.; Li, J.B.; Tsai, P.H.; Jang, J.S.C.; Huang, C.H.; Huang, J.C.; Huang, Y.S.; Lin, C.H.; et al. Synthesis and Characterization of an Open-Pore Toxic-Element-Free Ti-Based Bulk Metallic Glass Foam for Bio-Implant Application. J. Mater. Res. Technol. 2020, 9, 4518–4526.
  120. Meagher, P.; O’Cearbhaill, E.D.; Byrne, J.H.; Browne, D.J. Bulk Metallic Glasses for Implantable Medical Devices and Surgical Tools. Adv. Mater. 2016, 28, 5755–5762.
  121. Jin, C.; Liu, Z.; Yu, W.; Qin, C.; Yu, H.; Wang, Z. Biodegradable Mg–Zn–Ca-Based Metallic Glasses. Materials 2022, 15, 2172.
  122. Meng, D.; Yi, J.; Zhao, D.Q.; Ding, D.W.; Bai, H.Y.; Pan, M.X.; Wang, W.H. Tantalum Based Bulk Metallic Glasses. J. Non Cryst. Solids 2011, 357, 1787–1790.
  123. Chen, P.S.; Chen, H.W.; Duh, J.G.; Lee, J.W.; Jang, J.S.C. Characterization of Mechanical Properties and Adhesion of Ta-Zr-Cu-Al-Ag Thin Film Metallic Glasses. Surf. Coat. Technol. 2013, 231, 332–336.
  124. 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.
  125. Wataha, J.C.; Shor, K. Palladium Alloys for Biomedical Devices. Expert Rev. Med. Devices 2010, 7, 489–501.
  126. Woodward, B. Palladium in Temporary and Permanently Implantable Medical Devices. Platin. Met. Rev. 2012, 56, 213–217.
  127. Hua, N.; Hong, X.; Liao, Z.; Wang, Q.; Zhang, L.; Guo, Q.; Ye, X.; Brechtl, J.; Liaw, P.K. A Biocompatible Pd-Based BMG with Excellent Corrosive-Wear Resistance for Implant Applications. Intermetallics 2020, 124, 106847.
  128. Cihova, M.; Müller, E.; Chandorkar, Y.; Thorwarth, K.; Fortunato, G.; Maniura-Weber, K.; Löffler, J.F.; Rottmar, M. Palladium-Based Metallic Glass with High Thrombogenic Resistance for Blood-Contacting Medical Devices. Adv. Funct. Mater. 2022, 32, 2108256.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , ,
View Times: 605
Revisions: 2 times (View History)
Update Date: 28 Nov 2022
1000/1000
Video Production Service