Next Article in Journal
Double Perovskite Ba2LaTaO6 for Ultrafast Fiber Lasers in Anomalous and Normal Net Dispersion Regime
Next Article in Special Issue
Silver and Copper Nanoparticles Induce Oxidative Stress in Bacteria and Mammalian Cells
Previous Article in Journal
Artificial Synapse Consisted of TiSbTe/SiCx:H Memristor with Ultra-high Uniformity for Neuromorphic Computing
Previous Article in Special Issue
A Critical Review of the Antimicrobial and Antibiofilm Activities of Green-Synthesized Plant-Based Metallic Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Progress on Nanocrystalline Metallic Materials for Biomedical Applications

1
School of Materials Science and Engineering, University of Science and Technology Beijing, Beijing 100083, China
2
School of Engineering, RMIT University, Melbourne, VIC 3001, Australia
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(12), 2111; https://doi.org/10.3390/nano12122111
Submission received: 28 March 2022 / Revised: 25 May 2022 / Accepted: 30 May 2022 / Published: 19 June 2022

Abstract

:
Nanocrystalline (NC) metallic materials have better mechanical properties, corrosion behavior and biocompatibility compared with their coarse-grained (CG) counterparts. Recently, nanocrystalline metallic materials are receiving increasing attention for biomedical applications. In this review, we have summarized the mechanical properties, corrosion behavior, biocompatibility, and clinical applications of different types of NC metallic materials. Nanocrystalline materials, such as Ti and Ti alloys, shape memory alloys (SMAs), stainless steels (SS), and biodegradable Fe and Mg alloys prepared by high-pressure torsion, equiangular extrusion techniques, etc., have better mechanical properties, superior corrosion resistance and biocompatibility properties due to their special nanostructures. Moreover, future research directions of NC metallic materials are elaborated. This review can provide guidance and reference for future research on nanocrystalline metallic materials for biomedical applications.

1. Introduction

Biomedical metallic materials have good mechanical properties, biocompatibility, etc. However, the properties of conventional coarse-grained (CG) metallic materials do not fully meet clinical requirements. Nanocrystalline (NC) metallic materials have substantially improved mechanical properties, corrosion behavior, and biocompatibility than conventional CG metallic materials [1,2], thus it is obvious that the properties of nanocrystalline metallic materials can better meet clinical needs compared with conventional coarse-grained metallic materials. Therefore, NC metallic biomaterials are receiving increasing attention in recent years [3]. NC metallic biomaterials can be divided into two categories: bioinert NC metallic materials and biodegradable NC metallic materials. Currently, interest in bioinert NC metallic materials include NC pure titanium (Ti) and its alloys [4,5,6], NC SMAs [7], NC SS [8], and NC biodegradable metallic alloys including Fe-based alloys and Mg-based alloys [9]. The nanostructured metallic alloys open new avenues and concepts for medical implants, providing benefits in all areas of medical device technology.
Data from several studies suggest that after NC processing, NC metallic biomaterials have desirable characteristics, such as good manufacturability, and superior physical and mechanical properties [10,11]. This review summarizes the mechanical properties, corrosion behavior, biocompatibility, and clinical applications of bioinert NC metallic materials and biodegradable NC metallic materials in recent years and provides future directions for the development of biomedical NC metallic materials.

2. Bio-Inert NC Metallic Materials

2.1. Biomedical NC Pure Ti and Its Alloys

The excellent biocompatibility of medical devices manufactured from conventional CG pure Ti and its alloys for clinical applications has been demonstrated in several clinical investigations [12,13,14]. However, the mechanical strength and hardness of commercial CG Ti and its alloys need to be further improved. Nano-structuring of pure Ti and its alloys have recently been demonstrated to be a new horizon and promising alternative method for increasing the mechanical characteristics of conventional CG Ti alloys.
Researchers find that the mechanical performance of NC pure Ti is superior to that of conventional commercial CG Ti [15]. The strength of the NC Ti is nearly twice than that of conventional CG Ti: the σYS and σUTS of CG pure Ti were tested to be about 530 MPa and 700 MPa, respectively; after equal-channel angular pressing (ECAP), the CG Ti transformed to NC Ti and showed much higher σYS of 1267 MPa and σUTS of 1330 MPa [16]. In another study, the NC Ti-15Mo alloy showed significantly enhanced elastic modulus, microhardness, and tribological properties compared to its CG counterpart because of its higher relative density, sealed porosity, and grain size refinement. The binary NC Ti-15Mo alloy exhibited a microhardness of 315 HV0.02 and a modulus of elasticity of 95 GPa at 1373 K [17]. NC Ti-20Nb-13Zr has an average grain size of 70–140 nm with a duplex microstructure of the α-Ti (hcp) region surrounding the β-Ti (bcc) matrix, leading to a hardness of 660 HV, and the NC alloy also showed stimulation of new bone formation [18]. The σYS and σUTS of the NC Ti-5Ta-1.8Nb were reported to be 800 MPa and 750 MPa, respectively, and the fracture surface of the NC alloy exhibited shear bands and more ductile dimples compared to its CG counterpart which showed dimples and microvoids [19].
Not only have the strength and hardness of the NC Ti alloys increased, but their elongation has also been improved considerably compared to their CG counterparts. Shahmir et al. [20] obtained a good combination of strength and elongation in high-pressure torsion (HPT)-processed NC Ti at elevated temperatures ranging from 573 to 773 K. The strength, microhardness, and elongation of the NC pure Ti (Grade 2) with a grain size of 70 nm were measured as 945 MPa, 300 HV, and 9% at room temperature and approximately 350 MPa, 230 HV, and 130% at 673 K, respectively. Filho et al. [15] fabricated NC pure Ti (grade 2) with ECAP followed by cold rolling and found that this method gives the best strength-ductility combination. González-Masís et al. found that the nanocrystalline Ti-Nb-Zr-Ta-Hf had a combination of high hardness of 564 HV and moderate elastic of 79 GPa [21]. The mechanical properties of NC Ti better meet the requirements of bone replacement and repair. Two main factors can be identified to determine the strengthening mechanisms of NC metallic materials: grain refinement and increased dislocation density induced by the processing of the NC materials [22].
Elias et al. [23] compared the compressive and fatigue strength of dental implants made from the NC Ti (grade 4) processed by ECAP and the microcrystalline Ti (grade 4) and found that the NC Ti implants exhibited both higher compressive and fatigue strength for 5 × 106 cycles than the microcrystalline Ti implants, and the ECAP-processed NC Ti exhibited transgranular fracture with no striation at the fatigue crack initiation and propagation regions. The increased compression and fatigue strengths of the NC Ti make it a very good material for dental implant applications. Javadhesari et al. produced Ti-50 at%Cu alloy which showed excellent mechanical properties: ultra-high microhardness of 10 GPa and acceptable toughness of 8.14 MPa⋅m1/2 [24]. All the abovementioned studies have demonstrated that NC Ti and its alloys have better mechanical properties for implant applications.
In addition to the enhanced mechanical properties, an enhanced biological response can also be anticipated from NC metallic materials. For instance, the fibroblast mice cells L929 covered 53% commercially pure Ti (CP-Ti) surface and 87.2% nanostructured Ti surface after 72 h of cell culturing, indicating superior cytocompatibility of the NC Ti compared to its CG counterpart [4]. In another study, NC Ti showed an improvement in in vitro biosafety and long-term cellular functionalization in cytobiology and in vivo biostability [25]. Figure 1 shows the histotomy of bone contact of NC Ti at four weeks implantation as compared to CG Ti, indicating a higher osseointegration ability with freshly generated bone development direction and kinetics following implantation of NC Ti. Cell adhesion and proliferation test for nanocrystalline Ti25Nb16Hf showed lower adhesion and higher proliferation when compared to Ti grade 2 [26].
NC Ti-29Nb-13Ta-4.6Zr (the so-called TNTZ alloy) showed greater hardness than its CG counterpart, which is up to 310 HV [27]. Lin et al. [28] reported that the β Ti-35Nb-3Zr-2Ta alloy exhibited ultrafine equiaxed grains of approximately 300 nm after ECAP for 4 passes at 500 °C; the ECAP-processed alloy showed a longitudinal microhardness of 224 HV, a σYS of 390 MPa, and a σUTS of 765 MPa, while maintaining a good level of elongation of 16.5% and elastic modulus of 59 GPa. In another study, the microstructure of the HPT-processed TNTZ exhibited a single phase of β grains with diameters of a few hundred nanometers and high-angle boundaries, and due to the severe plastic deformation of the HPT process, the grains exhibited non-uniform subgrains with high dislocation density. The tensile strength of the nanocrystalline TNTZ alloy increased significantly [29]. In summary, the microhardness of NC TNTZ alloy is consistently much higher than its CG counterpart [30].
Xie et al. [31] reported that HPT-processed NC β-Ti alloy (Ti-36Nb-2.2Ta-3.7Zr-0.3O, at.%) showed markedly improved mechanical and biocompatibility properties; the hardness and elastic modulus of the NC Ti alloy were, respectively, 23% higher and 34% lower than those of its CG counterpart. The decrease in the elastic modulus of a metallic implant biomaterial is of critical significance because it helps prevent stress shielding that occurs when the implant is stiffer than its surrounding host bone. The β-type Ti-24Nb-4Zr-8Sn alloy processed by warm swaging and warm rolling with a uniform microstructure comprising a β phase with a size less than 200 nm and the precipitation of nanosized α phase, exhibited high ultimate tensile strength of 1150 MPa, low elastic modulus of 56 GPa, and good ductility with an elongation of 8%, along with large-scale nonlinear deformation behavior with a recoverable strain of up to 3.4% [32]. Kent et al. [33] reported the mechanical properties of the Ti-25Nb-3Zr-3Mo-2Sn alloy processed by a modified accumulative roll bonding (ARB) technique, and found that after 4 cycles of rolling, the ARB-processed sample exhibited significantly refined β grains heavily elongated in the rolling direction and NC α precipitates distributed on the β grain boundaries, with an ultimate tensile strength of 1220 MPa, a 0.5% proof stress of 946 MPa, which were, respectively, ~70% higher and almost double those of the CG solution treated counterpart. He et al. [34] investigated the mechanical properties of the Ti60Cu14Ni12Sn4M10 (M = Nb, Ta, Mo) alloys prepared using arc melting and copper mold casting. The alloys exhibited a composite microstructure containing a micrometer-sized dendritic β-Ti(M) phase dispersed in an NC matrix with a compressive elastic modulus in the range of 59–103 GPa, a compressive yield strength in the range of 1037–1755 MPa, and a compressive plastic strain in the range of 1.68–21.34% [35].
Some studies have also shown that NC Ti alloys have superior electrochemical corrosion resistance. Yilmazer et al. [36] evaluated the corrosion behavior of the HPT-processed Ti-29Nb-13Ta-4.6Zr alloy in simulated body fluid (SBF) using electrochemical impedance spectroscopy (EIS) and found that the NC alloy exhibited improved corrosion performance than its CG counterpart due to a passivated surface layer contained a titania (TiO2) matrix dispersed with zirconia (ZrO2), niobia (Nb2O5), and tantala (Ta2O5) oxides. Furthermore, an NC Ti alloy exhibits better hydrophilic property due to the NC structure [37]. This will lead to improved protein adsorption properties because of the increased contact points between the protein and the NC surface. The adsorption of bone morphogenetic proteins on a material surface affects the cell adhesion, spreading, and proliferation on the material. M. A. Hussein showed that nanocrystalline Ti20Nb20Zr alloy had a hydrophilic nature compared with a CP Ti [38]. Xie et al. [31] showed that HPT-processed NC β-type Ti-36Nb-2.2Ta-3.7Zr-0.3O (at.%) showed enhanced cell attachment and proliferation of human gingival fibroblasts (HGF) after 30 min cell seeding compared to its CG counterpart.

2.2. Biomedical NC SMAs

Nitinol refers to a family of SMAs composed of nickel (Ni) and Ti with a unique combination of properties including superelasticity and shape memory properties. The word nitinol is originated from its composition (Ni-Ti) and its place of discovery of Naval Ordnance Laboratory (USA) by William J. Buehler and Frederick Wang in 1963 [39]. Nitinol are the most widely used biomedical SMAs due to their large recoverable strains (~8%) in polycrystalline forms [40]. The phase transformation temperature of medical Ni-Ti SMAs is close to that of the human body; therefore, this class of Ti alloys plays an important role in the medical field that is unmatched by other materials. Conventionally, Ni-Ti SMAs are frequently used in orthopedics, dentistry, and cardiovascular stent treatments. NC Ni-Ti-based SMAs offer even better overall mechanical properties [41]. Ti-Zr-based alloys alloyed with Nb, Ta, Mo, and Sn are also another type of shape memory alloys. Sheremetyev et al. [42] reported that Ti-18Zr-15Nb alloy processed by ECAP at 250 °C for 7 passes showed a σYS of 962 MPa, a σUTS ultimate tensile strength of 988 MPa, and an elongation of 5.4%.
Yan et al. [43] found that compared to CG NiTi SMAs, the NC austenite NiTi (Ni-49.3Ti, at.%) SMA with a B2 (CsCl) type ordered structure showed a significantly enhanced compressive yield strength of 2552.1 MPa, and the value of stress-induced austenite transformation increased with decreasing grain size. Nie et al. [44] found that the HPT-processed NC Ni50.2Ti49.8 alloy exhibited a hardness of 456.8 ± 14.9 HV. Sharifi and Kermanpur [45] performed hot rolling and annealing at 900 °C on Ni50Ti50 alloy, followed by cold rolling with thickness reduction of 70% and annealing at 400 °C, and the resultant NC alloy exhibited superelastic properties including a recoverable strain of 12% and an upper plateau stress (σSIM) (i.e., the critical stress for stress-induced martensitic transformation) of 610 MPa, which is significantly higher than that of the CG alloy with a σSIM of 160 MPa. Baigonakova et al. [46] found that NiTi0.1Ag wires provided optimal strength (1450 MPa) and ductility (33.4%) properties, due to the dislocation-free homogeneous nanocrystalline structure.
However, NiTi SMAs contain a large amount of Ni ions. NiTi alloy implants might release Ni ions into the human tissue and cause severe allergic reactions after implantation in the body [47]. Nanocrystallization not only significantly improved the mechanical and superelastic properties of NiTi SMAs, but also fundamentally solved the problem of rapid Ni ion release because NC NiTi SMAs improved corrosion resistance compared to their CG counterparts. Shri et al. [48] found that through severe plastic deformation of HPT, the corrosion behavior of NiTi SMAs was changed by grain refinement, the NC Ti-50Ni (at.%) exhibited a stable, protective layer on its surface in a cell culture medium and increased corrosion resistance, leading to decreased Ni ion release. Nie et al. [44] reported superiorly higher corrosion resistance of HPT-processed NC Ni50.2Ti49.8 alloy with a substantially lower rate of Ni ion release than its microcrystalline counterpart in both Hanks’ solution and artificial saliva, which was far below the threatening threshold of a daily diet. The results of murine fibroblast (L-929) and osteoblast cell lines (MG63) cultured and cell proliferation are shown in Figure 2. There is no cytotoxicity for nanocrystalline Ni50.2Ti49.8 till 4 days culture with L-929 and MG63.
Li et al. [7] investigated the in vitro and in vivo biological properties of an ECAP-processed NC Ti49.2Ni50.8 alloy for orthopedic implant applications and indicated enhanced cell viability, adhesion, proliferation, ALP (alkaline phosphatase) activity, and mineralization than its CG counterpart.

2.3. Biomedical NC Stainless Steels (SSs)

Biomedical NC SSs have enhanced passivation behavior and corrosion resistance than their CG counterparts; furthermore, NC SSs also exhibit enhanced hydrophilic and protein adsorption properties, leading to improved biological properties including better cell attachment, spreading, and proliferation [49].
Previous studies have found that NC ASTM F-138 austenitic SSs have higher mechanical strength than their CG counterpart [15,50,51]. Heidari et al. found that the ASTM F2581 nanocrystalline stainless steels had 824 MPa yield strength, and the strength exceeded 1 GPa [51]. NC 304 SS fabricated by severe rolling showed a hardness of 480.0 ± 10.1 HV and improved corrosion resistance; Figure 3a shows the OCP curves of nanocrystalline 304 SS and microcrystalline 304 SS in artificial saliva. The result of OCP curve indicates that nanocrystalline stainless steels show more corrosion resistance than microcrystalline stainless steels in artificial saliva. Polarization studies (Figure 3b) revealed that NC 304 SS are more corrosion resistant than conventional CG 304 SS in an oral-like environment with higher corrosion potential [49].
Nie et al. using electrochemical measurement concluded a similar result that nanocrystalline stainless steel has higher corrosion resistance compared to microcrystalline stainless steel and the corrosion behavior of nanocrystalline 304 SS do not have significantly superior resistance to pitting corrosion compared to microcrystalline stainless steels [52].

2.4. Other Types of Bio-Inert NC Metallic Materials

There are also some other types of NC metallic materials for biomedical application. NC silver (Ag) prepared by spark plasma sintering (SPS) at 600 K for 5 min exhibited a mean grain size of 380 nm and showed a σYS 4.6 times higher than that of CG counterpart with a mean grain size of 49.65 µm and more than 30% uniform elongation [53]. The HPT-processed nanocrystalline CoCrMo exhibited improved tribocorrosion resistance but deteriorated corrosion resistance [54]. Huo et al. [55] fabricated an NC surface with an average grain size of ≤20 nm using sliding friction treatment (SFT) on CG pure Ta and comparatively studied the osteoblast cell responses to the CG and NC Ta using human osteoblastic hFOBl.19 cells. Their results showed that the NC surface exhibited higher surface hydrophilicity and enhanced corrosion resistance than the CG surface, thus leading to enhanced osteoblast adherence and spreading after 1 day’s cell culturing and markedly improved cell proliferation, maturation, and mineralization after 14 days’ cell culturing. Figure 4 shows the morphologies of hFOBl.19 cells on CG and NC pure Ta after 1, 3, and 7 days’ of culturing. The results of this study indicated the superior cytocompatibility of the NC Ta surface compared to its CG counterpart [55].
Some bio-inert NC metallic biomaterials published in the literature in recent years, their fabrication methods, grain sizes, mechanical properties, corrosion behaviors, biocompatibility, and potential applications are summarized in Table 1.

3. Biodegradable NC Metallic Materials for Biomedical Applications

Biodegradable metals have been research hotspots for the last two decades [62]. Nanocrystallization of biodegradable metals can further enhance their comprehensive properties. Wang et al. [63] performed hot rolling on Mg–2Zn alloy and found that the alloy exhibited a grain size of ~70 nm with a high σYS of 223 MPa and σUTS of 260 MPa, and a strong corrosion resistance with a corrosion rate of 0.2 mm/y in vivo when tested using Sprague Dawley rats. Nie et al. [9] demonstrated that ECAP-processed NC pure iron (Fe) exhibited higher corrosion resistance and improved hemocompatibility and cytocompatibility. Zhang et al. [64] investigated the corrosion resistance of an NC Mg-2Zn-0.24Ca (wt.%) alloy processed by HPT and annealing and reported that the grain size, number of (0002) oriented grains, second phase, and surface stress of the alloy changed with the annealing temperature and time, and these factors affected the corrosion rate. The HPT-processed alloy showed the best corrosion resistance with the maximum polarization resistance and lowest hydrogen evolution rate when annealed at 210 °C for 30 min.
The HPT-processed NC Mg-1Ca alloy showed at least 5-fold improvement in corrosion resistance than the CG alloy due to the separation of the second phase (Mg2Ca) particles and their continuous nanoprecipitation. Figure 5 shows the micrographs of CG and NC Mg-1Ca alloy. The microstructure of the homogenized CG Mg-1Ca alloy contained equiaxial grains (Figure 5a), while the NC alloy showed a much finer grain size of 100 ± 9 nm with a greater dislocation density and higher shear strength (Figure 5c) [65].
Gu et al. fabricated Mg-3Ca alloys with different grain sizes by melt-spinning at different wheel rotating speeds and their electrochemical test results showed that the corrosion rate of the NC Mg-3Ca in simulated body fluid was significantly reduced compared with CG Mg-3Ca, and the NC Mg-3Ca alloy showed a more uniform corrosion morphology. Further, the extract of the NC Mg-3Ca alloy showed no cytotoxicity in relation to L-929 cells, whereas the extract of CG Mg-3Ca alloy did. On the surface of the NC Mg-3Ca alloy, the L-929 cells exhibited improved adhesion than the CG Mg-3Ca alloy [66].
Table 2 summarizes some biodegradable NC metallic biomaterials published in the literature in recent years, their fabrication methods, grain sizes, and mechanical, corrosion, and biocompatibility properties. Most of the degradable nanocrystalline alloys reported thus far are Mg-based, with a small amount of Fe-based alloys; Zn-based alloys have not yet been publicly reported, and further development of Zn-based nanocrystalline alloys is needed in the future.

4. Conclusions

This article provides a review on the mechanical properties, corrosion behavior, biocompatibility, and clinical applications of different NC metallic biomaterials. The main conclusions are as follows:
  • Biomedical NC metallic materials, such as Ti and its alloys, TNTZ, NiTi SMAs, SS, and biodegradable Fe and Mg alloys have significantly improved tensile yield strength, ultimate tensile strength, and hardness without significant reduction in ductility.
  • Biomedical NC metallic materials, such as Ti and its alloys, TNTZ, NiTi SMAs, SS, and biodegradable Fe and Mg alloys have better corrosion resistance than their conventional CG metallic materials.
  • Nanocrystallization of metallic biomaterials can improve their biocompatibility due to the unique nanostructures on their surfaces.
  • In future research, the relationships between grain size, microstructural characteristics, and material properties of NC metallic materials should be systematically investigated. More research and development should be devoted to zinc-based degradable NC alloys, iron-based degradable NC alloys, and Mg and plural NC alloys.

Author Contributions

Writing—original draft preparation, H.L. and P.W.; writing—review and editing, H.L. and C.W.; funding acquisition, H.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the National Natural Science Foundation of China, China (grant No. 31700819), the Young Elite Scientists Sponsorship Program by CAST (YESS, grant No. 2018QNRC001) and the Fundamental Research Funds for the Central Universities and the Youth Teacher International Exchange and Growth Program (No. QNXM20210021).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gatina, S.; Semenova, I.; Leuthold, J.; Valiev, R. Nanostructuring and Phase Transformations in the β-Alloy Ti-15Mo during High-Pressure Torsion. Adv. Eng. Mater. 2015, 17, 1742–1747. [Google Scholar] [CrossRef]
  2. Liu, Y.; Li, K.; Luo, T.; Song, M.; Wu, H.; Xiao, J.; Tan, Y.; Cheng, M.; Chen, B.; Niu, X.; et al. Powder metallurgical low-modulus Ti–Mg alloys for biomedical applications. Mater. Sci. Eng. C 2015, 56, 241–250. [Google Scholar] [CrossRef]
  3. Pérez-Prado, M.; Gimazov, A.; Ruano, O.; Kassner, M.; Zhilyaev, A. Bulk nanocrystalline ω-Zr by high-pressure torsion. Scr. Mater. 2008, 58, 219–222. [Google Scholar] [CrossRef]
  4. Valiev, R.Z.; Semenova, I.P.; Latysh, V.V.; Rack, H.; Lowe, T.C.; Petruželka, J.; Dluhos, L.; Hrusak, D.; Sochova, J. Nanostructured Titanium for Biomedical Applications. Adv. Eng. Mater. 2008, 10, B15–B17. [Google Scholar] [CrossRef]
  5. Semenova, I.P.; Klevtsov, G.V.; Klevtsova, N.Y.A.; Dyakonov, G.S.; Matchin, A.A.; Valiev, R.Z. Nanostructured Titanium for Maxillofacial Mini-Implants.  Adv. Eng. Mater. 2016, 18, 1216–1224. [Google Scholar] [CrossRef]
  6. An, B.; Li, Z.; Diao, X.; Xin, H.; Zhang, Q.; Jia, X.; Wu, Y.; Li, K.; Guo, Y. In vitro and in vivo studies of ultrafine-grain Ti as dental implant material processed by ECAP. Mater. Sci. Eng. C 2016, 67, 34–41. [Google Scholar] [CrossRef]
  7. Li, H.; Nie, F.; Zheng, Y.; Cheng, Y.; Wei, S.; Valiev, R. Nanocrystalline Ti49.2Ni50.8 shape memory alloy as orthopaedic implant material with better performance. J. Mater. Sci. Technol. 2019, 35, 2156–2162. [Google Scholar] [CrossRef]
  8. Sheng, J.; Wei, J.; Li, Z.; Man, K.; Chen, W.; Ma, G.; Zheng, Y.; Zhan, F.; La, P.; Zhao, Y.; et al. Micro/nano-structure leads to super strength and excellent plasticity in nanostructured 304 stainless steel. J. Mater. Res. Technol. 2022, 17, 404–411. [Google Scholar] [CrossRef]
  9. Nie, F.L.; Zheng, Y.F.; Wei, S.C.; Hu, C.; Yang, G. In vitro corrosion, cytotoxicity and hemocompatibility of bulk nanocrystalline pure iron. Biomed. Mater. 2010, 5, 065015. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Wu, Q.; Miao, W.-S.; Zhang, Y.-D.; Gao, H.-J.; Hui, D. Mechanical properties of nanomaterials: A review. Nanotechnol. Rev. 2020, 9, 259–273. [Google Scholar] [CrossRef]
  11. Valiev, R.Z.; Prokofiev, E.A.; Kazarinov, N.A.; Raab, G.I.; Minasov, T.B.; Stráský, J. Developing Nanostructured Ti Alloys for Innovative Implantable Medical Devices. Materials 2020, 13, 967. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. McCracken, D.M. MSBE, Dental Implant Materials: Commerciallv Pure Titanium and Titanium Alloys. J. Prosthodont. 1999, 8, 40–43. [Google Scholar] [CrossRef]
  13. Dick, J.C.; Bourgeault, C.A. Notch Sensitivity of Titanium Alloy, Commercially Pure Titanium, and Stainless Steel Spinal Implants. Spine 2001, 26, 1668–1672. [Google Scholar] [CrossRef]
  14. Aparicio, F.J.G.C.; Fonseca, C.; Barbosa, M.; Planell, J.A. Corrosion behaviour of commercially pure titanium shot blasted with different materials and sizes of shot particles for dental implant applications. Biomaterials 2003, 24, 263–273. [Google Scholar] [CrossRef]
  15. Filho, A.D.A.M.; Sordi, V.L.; Kliauga, A.; Ferrante, M. The effect of equal channel angular pressing on the tensile properties and microstructure of two medical implant materials: ASTM F-138 austenitic steel and Grade 2 titanium. J. Phys. Conf. Ser. 2010, 240, 012130. [Google Scholar] [CrossRef] [Green Version]
  16. Polyakov, A.V.; Dluhoš, L.; Dyakonov, G.S.; Raab, G.I.; Valiev, R.Z. Recent Advances in Processing and Application of Nanostructured Titanium for Dental Implants. Adv. Eng. Mater. 2015, 17, 1869–1875. [Google Scholar] [CrossRef]
  17. Fellah, M.; Hezil, N.; Leila, D.; Samad, M.A.; Djellabi, R.; Kosman, S.; Montagne, A.; Iost, A.; Obrosov, A.; Weiss, S. Effect of sintering temperature on structure and tribological properties of nanostructured Ti–15Mo alloy for biomedical applications. Trans. Nonferrous Met. Soc. China 2019, 29, 2310–2320. [Google Scholar] [CrossRef]
  18. Hussein, M.A.; Suryanarayana, C.; Al-Aqeeli, N. Fabrication of nano-grained Ti–Nb–Zr biomaterials using spark plasma sin-tering. Mater. Des. 2015, 87, 693–700. [Google Scholar] [CrossRef]
  19. Bhaskar, P.; Dasgupta, A.; Sarma, V.S.; Mudali, U.K.; Saroja, S. Mechanical properties and corrosion behaviour of nanocrystalline Ti–5Ta–1.8Nb alloy produced by cryo-rolling. Mater. Sci. Eng. A 2014, 616, 71–77. [Google Scholar] [CrossRef]
  20. Shahmir, H.; Pereira, P.H.R.; Huang, Y.; Langdon, T.G. Mechanical properties and microstructural evolution of nanocrystalline titanium at elevated temperatures. Mater. Sci. Eng. A 2016, 669, 358–366. [Google Scholar] [CrossRef] [Green Version]
  21. González-Masís, J.; Cubero-Sesin, J.M.; Campos-Quirós, A.; Edalati, K. Synthesis of biocompatible high-entropy alloy TiNbZrTaHf by high-pressure torsion. Mater. Sci. Eng. A 2021, 825, 141869. [Google Scholar] [CrossRef]
  22. Purcek, G.; Yapici, G.G.; Karaman, I.; Maier, H.J. Effect of commercial purity levels on the mechanical properties of ultrafine-grained titanium. Mater. Sci. Eng. A 2011, 528, 2303–2308. [Google Scholar] [CrossRef]
  23. Elias, C.N.; Fernandes, D.; De Biasi, R.S. Comparative study of compressive and fatigue strength of dental implants made of nanocrystalline Ti Hard and microcrystalline Ti G4. Fatigue Fract. Eng. Mater. Struct. 2016, 40, 696–705. [Google Scholar] [CrossRef]
  24. Javadhesari, S.M.; Alipour, S.; Akbarpour, M. Biocompatibility, osseointegration, antibacterial and mechanical properties of nanocrystalline Ti-Cu alloy as a new orthopedic material. Colloids Surf. B Biointerfaces 2020, 189, 110889. [Google Scholar] [CrossRef]
  25. Nie, F.L.; Zheng, Y.F.; Wei, S.C.; Wang, D.S.; Yu, Z.T.; Salimgareeva, G.K.; Polyakov, A.V.; Valiev, R.Z. In vitro and in vivo studies on nanocrystalline Ti fabricated by equal channel angular pressing with microcrystalline CP Ti as control. J. Biomed. Mater. Res. Part A 2013, 101A, 1694–1707. [Google Scholar] [CrossRef] [PubMed]
  26. González, M.; Peña, J.; Gil, F.J.; Manero, J.M. Low modulus Ti–Nb–Hf alloy for biomedical applications. Mater. Sci. Eng. C 2014, 42, 691–695. [Google Scholar] [CrossRef]
  27. Yilmazer, H.; Niinomi, M.; Nakai, M.; Hieda, J.; Akahori, T.; Todaka, Y. Microstructure and Mechanical Properties of a Biomedical β-Type Titanium Alloy Subjected to Severe Plastic Deformation after Aging Treatment. Key Eng. Mater. 2012, 508, 152–160. [Google Scholar] [CrossRef]
  28. Lin, Z.; Wang, L.; Xue, X.; Lu, W.; Qin, J.; Zhang, D. Microstructure evolution and mechanical properties of a Ti–35Nb–3Zr–2Ta biomedical alloy processed by equal channel angular pressing (ECAP). Mater. Sci. Eng. C 2013, 33, 4551–4561. [Google Scholar] [CrossRef]
  29. Yilmazer, H.; Niinomi, M.; Akahori, T.; Nakai, M.; Todaka, Y. Effect of high-pressure torsion processing on microstructure and mechanical properties of a novel biomedical β-type Ti-29Nb-13Ta-4.6Zr after cold rolling. Int. J. Microstruct. Mater. Prop. 2012, 7, 168–186. [Google Scholar] [CrossRef]
  30. Yilmazer, H.; Niinomi, M.; Nakai, M.; Hieda, J.; Todaka, Y.; Akahori, T.; Miyazaki, T. Heterogeneous structure and mechanical hardness of biomedical β-type Ti–29Nb–13Ta–4.6Zr subjected to high-pressure torsion. J. Mech. Behav. Biomed. Mater. 2012, 10, 235–245. [Google Scholar] [CrossRef] [PubMed]
  31. Xie, K.Y.; Wang, Y.; Zhao, Y.; Chang, L.; Wang, G.; Chen, Z.; Cao, Y.; Liao, X.; Lavernia, E.J.; Valiev, R.Z.; et al. Nanocrystalline β-Ti alloy with high hardness, low Young’s modulus and excellent in vitro biocompatibility for bio-medical applications. Mater. Sci. Eng. C 2013, 33, 3530–3536. [Google Scholar] [CrossRef] [PubMed]
  32. Hao, Y.L.; Zhang, Z.B.; Li, S.J.; Yang, R. Microstructure and mechanical behavior of a Ti–24Nb–4Zr–8Sn alloy processed by warm swaging and warm rolling. Acta Mater. 2012, 60, 2169–2177. [Google Scholar] [CrossRef]
  33. Kent, D.; Wang, G.; Yu, Z.; Ma, X.; Dargusch, M. Strength enhancement of a biomedical titanium alloy through a modified ac-cumulative roll bonding technique. J. Mech. Behav. Biomed. Mater. 2011, 4, 405–416. [Google Scholar] [CrossRef] [PubMed]
  34. He, G.; Eckert, J.; Dai, Q.; Sui, M.; Löser, W.; Hagiwara, M.; Ma, E. Nanostructured Ti-based multi-component alloys with potential for biomedical applications. Biomaterials 2003, 24, 5115–5120. [Google Scholar] [CrossRef]
  35. He, G.; Hagiwara, M. Ti alloy design strategy for biomedical applications. Mater. Sci. Eng. C 2006, 26, 14–19. [Google Scholar] [CrossRef]
  36. Yilmazer, B.D.H.; Niinomi, M.; Nakai, M. Electrochemical Impedance Spectroscopy(EIS) Evaluation of Biomedical Nanostructured Β-Type Titanium Alloys. In Proceedings of the 1st International Symposium on Light Alloys and Composite Materials, Karabük, Turkey, 22–24 March 2018. [Google Scholar]
  37. Lu, J.; Zhang, Y.; Huo, W.; Zhang, W.; Zhao, Y.; Zhang, Y. Electrochemical corrosion characteristics and biocompatibility of nanostructured titanium for implants. Appl. Surf. Sci. 2018, 434, 63–72. [Google Scholar] [CrossRef]
  38. Hussein, M.A. Synthesis, Characterization, and Surface Analysis of Near-β Ti20Nb20Zr Alloy Proceeded by Powder Metallurgy for Biomedical Applications. JOM 2022, 74, 924–930. [Google Scholar] [CrossRef]
  39. Otsuka, K.; Ren, X. Physical metallurgy of Ti–Ni-based shape memory alloys. Prog. Mater. Sci. 2005, 50, 511–678. [Google Scholar] [CrossRef]
  40. Wen, C.E.; Xiong, J.Y.; Li, Y.C.; Hodgson, P.D. Porous shape memory alloy scaffolds for biomedical applications: A review. Phys. Scr. 2010, 2010, 014070. [Google Scholar] [CrossRef]
  41. Hao, S.; Cui, L.; Jiang, D.; Han, X.; Ren, Y.; Jiang, J.; Liu, Y.; Liu, Z.; Mao, S.; Wang, Y.; et al. A Transforming Metal Nanocomposite with Large Elastic Strain, Low Modulus, and High Strength. Science 2013, 339, 1191–1194. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Sheremetyev, V.; Churakova, A.; Derkach, M.; Gunderov, D.; Raab, G.; Prokoshkin, S. Effect of ECAP and annealing on structure and mechanical properties of metastable beta Ti-18Zr-15Nb (at.%) alloy. Mater. Lett. 2021, 305, 130760. [Google Scholar] [CrossRef]
  43. Yan, B.; Jiang, S.; Sun, D.; Wang, M.; Yu, J.; Zhang, Y. Martensite twin formation and mechanical properties of B2 austenite NiTi shape memory alloy undergoing severe plastic deformation and subsequent annealing. Mater. Charact. 2021, 178, 111273. [Google Scholar] [CrossRef]
  44. Nie, F.; Zheng, Y.; Cheng, Y.; Wei, S.; Valiev, R. In vitro corrosion and cytotoxicity on microcrystalline, nanocrystalline and amorphous NiTi alloy fabricated by high pressure torsion. Mater. Lett. 2010, 64, 983–986. [Google Scholar] [CrossRef]
  45. Sharifi, E.M.; Kermanpur, A. Superelastic properties of nanocrystalline NiTi shape memory alloy produced by thermomechanical processing. Trans. Nonferrous Met. Soc. China 2018, 28, 515–523. [Google Scholar] [CrossRef]
  46. Baigonakova, G.; Marchenko, E.; Chekalkin, T.; Kang, J.-H.; Weiss, S.; Obrosov, A. Influence of Silver Addition on Structure, Martensite Transformations and Mechanical Properties of TiNi–Ag Alloy Wires for Biomedical Application. Materials 2020, 13, 4721. [Google Scholar] [CrossRef] [PubMed]
  47. Ozan, S.; Munir, K.; Biesiekierski, A.; Ipek, R.; Li, Y.; Wen, C. 1.3.3A—Titanium Alloys, Including Nitinol. In Biomaterials Science, 4th ed.; Wagner, W.R., Sakiyama-Elbert, S.E., Zhang, G., Yaszemski, M.J., Eds.; Academic Press: Cambridge, MA, USA, 2020; pp. 229–247. [Google Scholar]
  48. Shri, D.N.A.; Tsuchiya, K.; Yamamoto, A. Corrosion Behavior of HPT-Deformed TiNi Alloys in Cell Culture Medium. AIP Conf. Proc. 2017, 1877, 030010. [Google Scholar] [CrossRef] [Green Version]
  49. Nie, F.; Wang, S.; Wang, Y.; Wei, S.; Zheng, Y. Comparative study on corrosion resistance and in vitro biocompatibility of bulk nanocrystalline and microcrystalline biomedical 304 stainless steel. Dent. Mater. 2011, 27, 677–683. [Google Scholar] [CrossRef]
  50. Yin, F.; Hu, S.; Xu, R.; Han, X.; Qian, D.; Wei, W.; Hua, L.; Zhao, K. Strain rate sensitivity of the ultrastrong gradient nanocrystalline 316L stainless steel and its rate-dependent modeling at nanoscale. Int. J. Plast. 2020, 129, 102696. [Google Scholar] [CrossRef]
  51. Heidari, L.; Hadianfard, M.; Khalifeh, A.; Vashaee, D.; Tayebi, L. Fabrication of nanocrystalline austenitic stainless steel with superior strength and ductility via binder assisted extrusion method. Powder Technol. 2020, 379, 38–48. [Google Scholar] [CrossRef]
  52. Nie, F.L.; Wang, Y.B.; Wei, S.C.; Zheng, Y.F.; Wang, S.G. In Vitro Corrosion and Haemocompatibility of Bulk Nanocrystalline 304 Stainless Steel by Severe Rolling. Mater. Sci. Forum 2010, 667–669, 1113–1118. [Google Scholar] [CrossRef] [Green Version]
  53. Mineta, T.; Saito, T.; Yoshihara, T.; Sato, H. Structure and mechanical properties of nanocrystalline silver prepared by spark plasma sintering. Mater. Sci. Eng. A 2019, 754, 258–264. [Google Scholar] [CrossRef]
  54. Namus, R.; Rainforth, W.M.; Huang, Y.; Langdon, T.G. Effect of grain size and crystallographic structure on the corrosion and tribocorrosion behaviour of a CoCrMo biomedical grade alloy in simulated body fluid. Wear 2021, 478–479, 203884. [Google Scholar] [CrossRef]
  55. Huo, W.T.; Zhao, L.Z.; Yu, S.; Yu, Z.T.; Zhang, P.X.; Zhang, Y.S. Significantly enhanced osteoblast response to nano-grained pure tantalum. Sci. Rep. 2017, 7, 40868. [Google Scholar] [CrossRef] [Green Version]
  56. Fouzia, H.; Mamoun, F.; Hezil, N.; Aissani, L.; Goussem, M.; Said, M.; Mohammed, A.S.; Montagne, A.; Iost, A.; Weiß, S.; et al. The Effect of Milling Time on the Microstructure and Mechanical Properties of Ti-6Al-4Fe Alloys. Mater. Today Commun. 2021, 27, 102428. [Google Scholar] [CrossRef]
  57. Klinge, L.; Siemers, C.; Sobotta, B.; Krempin, H.; Kluy, L.; Groche, P. Nanocrystalline Ti13Nb13Zr for Dental Implant Applications. In Proceedings of the 60th Conference of Metallurgists, COM, Online, 17–19 August 2021. [Google Scholar]
  58. Aguilar, C.; Pio, E.; Medina, A.; Parra, C.; Mangalaraja, R.; Martin, P.; Alfonso, I.; Tello, K. Effect of Sn on synthesis of nanocrystalline Ti-based alloy with fcc structure. Trans. Nonferrous Met. Soc. China 2020, 30, 2119–2131. [Google Scholar] [CrossRef]
  59. Yilmazer, H.; Niinomi, H.M.; Nakai, M.; Cho, K.; Hieda, J.; Todaka, Y.; Miyazaki, T. Mechanical properties of a medical β-type titanium alloy with specific microstructural evolution through high-pressure torsion. Mater. Sci. Eng. C 2013, 33, 2499–2507. [Google Scholar] [CrossRef] [PubMed]
  60. Despang, F.; Bernhardt, A.; Lode, A.; Hanke, T.; Handtrack, D.; Kieback, B.; Gelinsky, M. Response of human bone marrow stromal cells to a novel ultra-fine-grained and dispersion-strengthened titanium-based material. Acta Biomater. 2010, 6, 1006–1013. [Google Scholar] [CrossRef] [PubMed]
  61. Huang, C.; Gao, Y.; Yang, G.; Wu, S.; Li, G.; Li, S. Bulk nanocrystalline stainless steel fabricated by equal channel angular pressing. J. Mater. Res. 2006, 21, 1687–1692. [Google Scholar] [CrossRef]
  62. Liu, H.-N.; Zhang, K.; Li, X.-G.; Li, Y.-J.; Ma, M.-L.; Shi, G.-L.; Yuan, J.-W.; Wang, K.-K. Microstructure and corrosion resistance of bone-implanted Mg–Zn–Ca–Sr alloy under different cooling methods. Rare Met. 2021, 40, 643–650. [Google Scholar] [CrossRef]
  63. Wang, W.; Blawert, C.; Zan, R.; Sun, Y.; Peng, H.; Ni, J.; Han, P.; Suo, T.; Song, Y.; Zhang, S.; et al. A novel lean alloy of biodegradable Mg–2Zn with nanograins. Bioact. Mater. 2021, 6, 4333–4341. [Google Scholar] [CrossRef]
  64. Zhang, C.; Guan, S.; Wang, L.; Zhu, S.; Chang, L. The microstructure and corrosion resistance of biological Mg–Zn–Ca alloy processed by high-pressure torsion and subsequently annealing. J. Mater. Res. 2017, 32, 1061–1072. [Google Scholar] [CrossRef]
  65. Parfenov, E.; Kulyasova, O.; Mukaeva, V.; Mingo, B.; Farrakhov, R.; Cherneikina, Y.; Yerokhin, A.; Zheng, Y.; Valiev, R. Influence of ultra-fine grain structure on corrosion behaviour of biodegradable Mg-1Ca alloy. Corros. Sci. 2019, 163, 108303. [Google Scholar] [CrossRef]
  66. Gu, X.N.; Li, X.L.; Zhou, W.R.; Cheng, Y.; Zheng, Y.F. Microstructure, biocorrosion and cytotoxicity evaluations of rapid solidified Mg–3Ca alloy ribbons as a biodegradable material. Biomed. Mater. 2010, 5, 035013. [Google Scholar] [CrossRef] [Green Version]
  67. Kowalski, M.J.K. Ultrafine grained Mg-1Zn-1Mn-0.3Zr alloy and its corrosion behaviour. J. Achiev. Mater. Manuf. Eng. 2016, 74, 53–59. [Google Scholar] [CrossRef]
Figure 1. Histotomy of bone contact of: (a) CG Ti, and (b) NC Ti at four weeks, illustrated by fluorescence-dyeing reagents. Reprinted with permission from Ref. [25]. Copyright 2012, John Wiley and Sons.
Figure 1. Histotomy of bone contact of: (a) CG Ti, and (b) NC Ti at four weeks, illustrated by fluorescence-dyeing reagents. Reprinted with permission from Ref. [25]. Copyright 2012, John Wiley and Sons.
Nanomaterials 12 02111 g001
Figure 2. Cytotoxicity of (a) L-929 and (b) MG63 cell lines co-cultured with extracts from microcrystalline Ni50.2Ti49.8, and nanocrystalline Ni50.2Ti49.8. Reprinted with permission from Ref. [44]. Copyright 2010, Elsevier.
Figure 2. Cytotoxicity of (a) L-929 and (b) MG63 cell lines co-cultured with extracts from microcrystalline Ni50.2Ti49.8, and nanocrystalline Ni50.2Ti49.8. Reprinted with permission from Ref. [44]. Copyright 2010, Elsevier.
Nanomaterials 12 02111 g002
Figure 3. The electrochemical curves of (a) OCP and (b) polarization of microcrystalline 304 SS and nanocrystalline 304 SS in artificial saliva. Reprinted with permission from Ref. [49]. Copyright 2011, Elsevier.
Figure 3. The electrochemical curves of (a) OCP and (b) polarization of microcrystalline 304 SS and nanocrystalline 304 SS in artificial saliva. Reprinted with permission from Ref. [49]. Copyright 2011, Elsevier.
Nanomaterials 12 02111 g003
Figure 4. Typical morphologies of hFOBl.19 cells cultured on: (a,c,e) CG and (b,d,f) NC Ta surfaces for (a,b) 1 day, (c,d) 3 days, and (e,f) 7 days. Arrows indicate filopodia extensions. Reprinted with permission from Ref. [55].
Figure 4. Typical morphologies of hFOBl.19 cells cultured on: (a,c,e) CG and (b,d,f) NC Ta surfaces for (a,b) 1 day, (c,d) 3 days, and (e,f) 7 days. Arrows indicate filopodia extensions. Reprinted with permission from Ref. [55].
Nanomaterials 12 02111 g004
Figure 5. Microstructures of Mg-1Ca alloy: (a,b) scanning electron microscopy images of CG alloy; (c,d) transmission electron microscopy images of NC alloy. Reprinted with permission from Ref. [65].Copyright 2020, Elsevier.
Figure 5. Microstructures of Mg-1Ca alloy: (a,b) scanning electron microscopy images of CG alloy; (c,d) transmission electron microscopy images of NC alloy. Reprinted with permission from Ref. [65].Copyright 2020, Elsevier.
Nanomaterials 12 02111 g005
Table 1. Summary of bio-inert nanocrystalline metallic biomaterials.
Table 1. Summary of bio-inert nanocrystalline metallic biomaterials.
MaterialsMethodsGrain Size (nm)Mechanical PropertiesCorrosion PropertiesBiocompatibilityApplicationRef.
Pure Ti (grade 2)ECAP + cold rollingAbout 500σYS: Increased to 796 MPa
σUTS: Increased to 876 MPa
Hardness: Increased to 271 HV
--Bone
replacement and repair
[15]
Pure Ti (grade 4)ECAP150Good plasticity under compression
Fatigue strength higher than conventional CG Ti G4
--Dental
implant
[23]
Pure Ti (grade 4)SPD+TMT150σUTS: 1240 MPa
σYS: 1200 MPa
Elongation: 12%
Fatigue strength at 106 cycles: 620 MPa
-(Fibroblast mice cells L929)
Occupied surface after 72 h conventional
CP Ti: 53.0%
NC CP grade 4: 87.2%
Dental
implant
[4]
CP Ti (grade 4)ECAP250Ductility: 5%
TS: 1190 MPa
σUTS: 1240 MPa
Elongation: 11.5%
Fatigue strength at 106 cycles: 620 MPa
-Protein adsorption: Better than CP Ti
L929: Cell viability increases and growth ahead
MG63: Grow well
VSMCs: Not proliferate well
ECV304: Grow well
Hemolysis rates: Less than 5%
In vivo test: Superior bone formation
Bone
replacement
[25]
Pure Ti (grade 2)HPT70Strength: 940 MPa
High hardness: 300 HV
Elongation: 130%
---[20]
Pure Ti (grade 2)ECAE300σYS: 620 MPa
Ductility: 21%
---[22]
Pure Ti (grade 4)ECAE300σYS: 758 MPa
Ductility: 25%
---[22]
Pure TiECAP-conform + Drawing-σUTS: 1330 MPa
σYS: 1267 MPa
Elongation: 11%
Endurance limit: 107 cycles of 620 MPa
--Dental
implants
[16]
B2 austenite NiTi shape memory alloy (Ni-49.3 at.%Ti)SPD. + Annealing 4 h45Compressive yield stress: 2552.1 MPa
Fracture strain decreased 11.7%
σSIM: 267.8 MPa
---[43]
Ti-15MoHigh energy ball mill + Hot isostaticallypressed29 (1373 K)Microhardness: 315 HV0.02 (1373 K)
Elastic modulus: 95 GPa
Friction coefficient: 023–0.35 (1373 K)
---[17]
Ti-50at.%NiHPT--Increase corrosion resistance in the cell culture medium (stable and protective passive film)--[48]
Ni50Ti5070% cold rolling + annealing at 400 °C for 1 h20–70σSIM: 610 MPa---[45]
Ni50.2Ti49.8HPT-Hardness: 456.8 ± 14.9 HVSuperiorly higher corrosion
resistance than
microcrystalline Ni50.2Ti49.8 (Hanks’ solution and artificial saliva)
L-929: No cytotoxicity
MG63: No cytotoxicity
-[44]
Ti49.2Ni50.8ECAP150–250--Hemolysis rates: 0.1%
Number of adhered platelets: Lower than microcrystalline
Cell viability: Higher
Better osteogenesis functions
In vivo: Enhanced cell viability,
adhesion, proliferation, ALP activity, and
mineralization, and increased periphery
thickness of new bone
Orthopedic biomaterials[7]
Ti-6Al-4FeMechanical alloying-Hardness: 335 ± 17 HV0.05 (powders milled for 2 h), 387 ± 19 HV0.05 (powders milled for 6 h), 475 ± 23HV0.05 (powders milled for 12 h), 660 ± 33 HV0.05 (powders milled for 18 h)
Young’s modulus: 110–197 GPa
--Bone
replacement
[56]
Ti-5Ta-1.8NbCryo-rolling20σYS: About 800 MPa
σUTS: ~750 MPa
Elongation: About 5%
---[19]
Ti13Nb13ZrSPD200Young’s modulus: 72 GPa
σYS: 1150 MPa
Hardness: 300 HV
--Dental
implant
[57]
Ti-18Zr-15NbECAP at 250 °C for 7 passes20–100σYS: 962 MPa
σUTS: 988 MPa
Ductility: 5.4%
---[42]
Ti-20Nb-13ZrSPS-Hardness: 660 HV-Stimulate new bone formationDental and orthopedic applications[18]
Ti-13Ta-xSn (x = 3, 6, 9 and 12, at.%)Mechanical alloying10----[58]
Ti25Nd16HfCold rolling at 95% reduction50Ductility: 4.0%
σYS: 790 MPa
σUTS: 870 MPa
Elastic modulus: 42.3 GPa
The highest corrosion resistance
(corrosion current density 1.52 μA/cm2)
compared with Ti25Nb16Hf (0% C.R.) and Ti grade II
Cytotoxicity: ExcellentCell
adhesion (MG63 cells): Lower than pure Ti
Cell proliferation: Properly
-[26]
TiNbZrTaHfHPT<100Hardness: 564 ± 22 HV
Elastic modulus: 79 ± 3 GPa
Good plasticity under localized compression
---[21]
Ti-29Nb-13Ta-4.6ZrHPT40–500Ductility: Decrease
σUTS: Increase
Hardness: Great
---[29]
Ti-35Nb-3Zr-2TaECAP300–600Ductility: 16%
σUTS: 765 MPa
Elastic modulus: 59GPa
---[28]
Ti-24Nb-4Zr-8SnWarm swaging and warm rolling-Recoverable strain: 3.4%σUTS: 1150 MPaElastic modulus:56 GPaDuctility: 8%---[32]
Ti-29Nb-13Ta-4.6ZrHPT-σUTS: 800–1100 MPa
Elongation: 7%
Young’s modulus: 60 MPa
---[59]
Ti-29Nb-13Ta-4.6Zr (TNTZ)HPT-Hardness: Higher than 310 HV---[27]
Ti-29Nb-13Ta-4.6ZrHPT-Hardness: >183 HV
(Hardness values of peripheral region higher than that of central region)
---[30]
Ti-36Nb-2.2Ta-3.7Zr-0.3OHPT-Elastic modulus: 43 GPa (30% lower than CG counterpart)
Hardness: 320HV (23% higher than CG counterpart)
-Human gingival fibroblasts:
Attachment and proliferation were enhanced
Human dental follicular
Cells: Higher cell density
-[31]
Ti-25Nb-3Zr-3Mo-2SnAccumulative roll bonding130σUTS: 1220 MPa
0.5% proof stress: 946 MPa
Ductility: 4.5%
---[33]
Ti60Cu14Ni12Sn4Nb10Arc melting and copper mold casting-σYS: 1052 MPa
Young’s modulus: 59 GPa
Strain at yield point: 2.1%
---[35]
Ti 60Cu14Ni12Sn4M10 (M = Nb, Ta, Mo)Arc melting and copper mold casting-σYS: 1037–1755 MPa
Young’s modulus: 59–103 GPa
Plastic strain: Up to 21%
---[34]
Ti/1.3HMDSPowder metallurgy365Hardness: 320 HVYoung’s modulus: 129 MPa
σYS: 1439 MPa
Breaking elongation: 7.1%
Osteogenically induced hMSC: Comparable with CP Ti and Ti6Al4V-Bone repair[60]
ASTM F-138 austenitic steelECAP + cold rolling100–200YS: Increased to 1055 MPa
σUTS: Increased to 1059 MPa
Hardness: Increased to 339 HV
--Bone
replacement and repair
[15]
Ti-CuMechanical alloying and sintering Hardness: 10 GPaThe corrosion behavior of the alloy was slightly lower than cp-Ti98% anti-bacterial rate against Staphylococcus aureus (S. aureus) and Escherichia coli (E. coli), excellent cell viability to MG-63 cells, and high osteoblast formation rateOrthopedic material[24]
304 stainless steelSevere rolling-Strength 1280 MPa (NC 304), 640 MPa (CG 304 SS)More corrosion
resistant than the microcrystalline 304 SS in artificial saliva
Cytotoxicity (murine
fibroblast cells): Better than microcrystalline 304 SS
-[49]
304 stainless steelSevere rolling50Hardness: 480.0 ± 10.1 HVBetter corrosion
resistance (Hanks’ solution)
Cytotoxicity (L-929, NIH 3T3): No toxic effect,
Low hemolysis rate
-[52]
316L stainless steelSevereplastic deformationAround 5 at the surfaceMaximum nanohardness: 6.2 GPa
Young’s modulus: 200–220 GPa
---[50]
Stainless steelECAP74 (strain-induced martensite, BCC); 31 (austenite, FCC)----[61]
Austenitic stainless steelBinder assisted extrusion-Compressive yield strength: 824 MPa
Compressive strength: 1326 MPa
Uniform elongation: 50%
Hardness: 339 HV
---[51]
Pure silverSpark plasma sintering (sintered 600 K for 5 min)300σYS: 146 MPa
Uniform elongation: 30%
---[53]
CoCrMoFive-turns HPT-Compressive yield strength: 1.25 GPa
Hardness: 9.3 GPa
Elasticity modulus: 203 GPa
Reduce corrosion resistanceImproved tribocorrosion resistanceHip and knee replacements[54]
Table 2. Summary of biodegradable nanocrystalline metallic materials.
Table 2. Summary of biodegradable nanocrystalline metallic materials.
MaterialsMethodsGrain Size (nm)Mechanical PropertiesCorrosion PropertiesBiocompatibilityRef.
Pure FeECAP
(8 passes)
-σUTS: 470 MPa (double of CG counterpart)
Hardness: 444 ± 31 kg f mm−2 (4 times of CG counterpart)
Higher corrosion resistanceBetter hemocompatibility: Hemolysis less than 5%
VSMCs: Inhibited (less than 60%)
ECs and L929: Improved (more than 80%)
[9]
Mg-2wt.%ZnHot-rolled70σYS: 223 MPa
σUTS: 260 MPa
Good corrosion resistance (corrosion rate in vivo: 0.2 mm/y)-[63]
Mg-1Zn-1Mn-0.3Zr20 h Ball milling + annealing45-Corrosion resistance in Ringer solution improved-[67]
Mg-Zn-CaHPTWith an increase in annealing temperature, grain size increased from 100 to 900 Corrosion resistance increases with annealing temperature increased from 90–210 °C
Corrosion resistance decreases with temperature increased after 210 °C
-[64]
Mg-1CaHPT +
Annealing
100-Increased corrosion resistance-[65]
Mg-3CaMelt-spinning200–500-Uniform corrosion morphologyNo toxicity and improved adhesion in relation to L-929 cells[66]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, H.; Wang, P.; Wen, C. Recent Progress on Nanocrystalline Metallic Materials for Biomedical Applications. Nanomaterials 2022, 12, 2111. https://doi.org/10.3390/nano12122111

AMA Style

Li H, Wang P, Wen C. Recent Progress on Nanocrystalline Metallic Materials for Biomedical Applications. Nanomaterials. 2022; 12(12):2111. https://doi.org/10.3390/nano12122111

Chicago/Turabian Style

Li, Huafang, Pengyu Wang, and Cuie Wen. 2022. "Recent Progress on Nanocrystalline Metallic Materials for Biomedical Applications" Nanomaterials 12, no. 12: 2111. https://doi.org/10.3390/nano12122111

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