Next Article in Journal
Advanced Cold-Spraying Technology
Next Article in Special Issue
Complex Bioactive Chitosan–Bioglass Coatings on a New Advanced TiTaZrAg Medium–High-Entropy Alloy
Previous Article in Journal
The Role of Biochar Nanoparticles Performing as Nanocarriers for Fertilizers on the Growth Promotion of Chinese Cabbage (Brassica rapa (Pekinensis Group))
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mechanical and Anticorrosive Properties of TiNbTa and TiNbTaZr Films on Ti-6Al-4V Alloy

1
Department of Optoelectronics and Materials Technology, National Taiwan Ocean University, Keelung 202301, Taiwan
2
Center of Excellence for Ocean Engineering, National Taiwan Ocean University, Keelung 202301, Taiwan
3
Department of Materials Engineering, Ming Chi University of Technology, New Taipei 243303, Taiwan
4
Center for Plasma and Thin Film Technologies, Ming Chi University of Technology, New Taipei 243303, Taiwan
*
Author to whom correspondence should be addressed.
Coatings 2022, 12(12), 1985; https://doi.org/10.3390/coatings12121985
Submission received: 17 November 2022 / Revised: 14 December 2022 / Accepted: 15 December 2022 / Published: 18 December 2022

Abstract

:
In this study, TiNbTa and TiNbTaZr films were utilized as protective coatings on a Ti-6Al-4V alloy to inhibit corrosive attacks from NaCl aqueous solution and simulated body fluid. The structural and mechanical properties of multicomponent TiNbTa(Zr) films were investigated. The corrosion resistance of the TiNbTa(Zr)-film-modified Ti-6Al-4V alloy was evaluated using potentiodynamic polarization tests in a NaCl aqueous solution. The results indicate that the TiNbTa(Zr) films with high Ti and Zr contents exhibited inferior corrosive resistance related to the films with high Ta and Nb contents. Moreover, the TiNbTa(Zr)-coated Ti-6Al-4V plates were immersed in Ringer’s solution for eight weeks; this solution was widely used as a simulated body fluid. The formation of surficial oxide layers above the TiNbTa(Zr) films was examined using transmission electron microscopy and X-ray photoelectron spectroscopy, which prevented the elution of Al and V from the Ti-6Al-4V alloy. Ti33Nb19Ta21Zr27, Ti15Nb68Ta8Zr9, and Ti8Nb8Ta79Zr5 films are suggested as preferential candidates for TiNbTa(Zr)/Ti-6Al-4V assemblies applied as biocompatible materials.

1. Introduction

The development of biocompatible long-term implants without evident side effects has attracted the interest of researchers [1,2]. The literature indicates that high strength, high-wear resistance, and low Young’s modulus are the crucial requirements for orthopedic long-term implantation [3,4]. Among these material characteristics, low Young’s modulus is the essential property preventing the “stress shielding” effect [5]. In contrast to periprosthetic bone with a Young’s modulus less than 30 GPa [6], the stiffer artificial implants endure the majority of the load, resulting in bone resorption accompanied with porous constructions [7,8,9]. Ti-base alloys with a low Young’s modulus were utilized as implants [6]. Ti-6Al-4V has been widely used in medical applications due to its excellent mechanical properties, corrosion resistance, and biocompatibility [10,11,12]. However, the eluted toxic ions, aluminum and vanadium, could cause undesirable neurological disorders, Alzheimer’s disease, and local inflammation in the biological system [13], and the elution behavior needs to be inhibited. Ti-based alloys comprise two distinct crystalline structures, α-phase and β-phase, which are hexagonal close-packed (hcp) and body-centered cubic (bcc) phases [14], respectively. Nb, Ta, and Mo are recognized as β-phase stabilizers for Ti-alloys and β-phase Ti-alloys exhibit lower Young’s modulus and higher strength related to those of α-phase Ti-alloys [14,15,16]. TiNbTa [3], TiNbZrTa, and TiMoZrTa [14] alloys have been proposed as alternatives for biocompatible materials. Ti, Nb, and Ta elements, as well as Zr, Ru, Sn, and Au, exhibit exceptional biocompatibility in the human body [17]. Moreover, TiNbTa alloys have been shown to exhibit excellent corrosion resistance in contrast to the Ti-6Al-4V alloy due to the formation of a passivation oxide film consisting of Nb2O5 and Ta2O5; therefore, TiNbTa alloys without the elution effect of Al and V ions have been proposed as candidates for orthopedic implants [3].
Combining TiTaNb films and a commercial Ti-6Al-4V alloy has been recommended for application as implants [18]; however, the number of investigations into the aforementioned combination is limited. Lai et al. [19] have reported that TiNbZr and TiNbZrTa films fabricated through cathodic arc evaporation assisted the survival and growth of cells and were applied for orthopedic or dental implants. TiTaHfNbZr high-entropy alloy films prepared on Ti-6Al-4V have been explored [20,21], and these TiTaHfNbZr films exhibited an amorphous structure. This study aimed to evaluate the feasibility of TiNbTa(Zr)/Ti-6Al-4V assembly for implant materials. The low- and medium-entropy TiNbTa(Zr) films were fabricated through co-sputtering. The chemical compositions of TiNbTa(Zr) films were adjusted by distinct sputter powers applied to these targets. The phase structures and mechanical properties of the TiNbTa(Zr) films were studied. The elemental effects on the corrosion resistance of TiNbTa(Zr) films in NaCl solutions and the stability of TiNbTa(Zr) films in Ringer’s solution were explored.

2. Materials and Methods

TiNbTa(Zr) films were deposited on p-type (100) silicon wafers and Ti-6Al-4V coupons through co-sputtering with targets of pure Ti, Nb, Ta, and Zr. The sputtering apparatus is described in [22]. The Ti-6Al-4V coupons with dimensions of 20 × 20 × 3 mm3 were grounded and polished before deposition. The substrates were sputtered with a Ti interlayer for enhancing the adhesion between the TiNbTa(Zr) films and substrates. Then, various powers presented in Table 1 were set for co-sputtering TiNbTa and TiNbTaZr films with sputtering times of 50 and 90 min, respectively. The base pressure of the sputtering chamber was below 4 ×10−5 Pa, whereas the working pressure was set at 1.6 ×10−1 Pa under a 20 sccm flow of pure Ar. The substrate holder was rotated at a speed of 30 rpm during the co-sputtering process in order to homogenize the films’ compositions. These films were further immersed in a Ringer’s solution, a simulated body fluid, at 37 °C for eight weeks to evaluate the elution test.
The chemical compositions of the TiNbTa(Zr) films were examined using a field-emission electron probe microanalyzer (JXA-iHP200F, JEOL, Akishima, Japan). The phase constitutions of the films were confirmed using an X-ray diffractometer (XRD, X’Pert PRO MPD, PANalytical, Almelo, The Netherlands) under the grazing incident mode. The lattice constants, a0, of cubic phases were evaluated according to the following equation [23]:
a = a 0 + K × cos 2 θ sin θ
where a is the lattice constant of the individual reflection, θ is the diffraction angle, and K is a constant. The grain sizes of the films were evaluated on the basis of a Bragg–Brentano scan (θ–2θ scan) mode by using the Scherrer formula [23]:
D = 0.9 λ β cos θ B
where D is the average grain size, β is the full width at half maximum (FWHM) of reflection, θB is the Bragg angle, and λ is the X-ray wavelength.
The nanostructures of these films with protective C and Pt layers, prepared using a focused ion beam system (NX2000, Hatachi, Tokyo, Japan), were observed by transmission electron microscopy (TEM, JEM-2010E, JEOL, Akishima, Japan). The hardness and Young’s modulus values of the films were measured using a nanoindentation tester (TI-900 Triboindenter, Hysitron, Minneapolis, MN, USA) with a depth of 70 nm. The corrosion behavior of the TiNbTa(Zr) films and bare Ti-6Al-4V coupons were evaluated via potentiodynamic polarization tests (SP-200, BioLogic, Seyssinet-Pariset, France) in a 3.5 wt.% NaCl aqueous solution within a potential range of −2.5 V to 2.5 V. The compositional depth profiles and bonding characteristics of the films and bare Ti-6Al-4V alloy after immersing in Ringer’s solution were analyzed using X-ray photoelectron spectroscopy (XPS; PHI 1600, PHI, Kanagawa, Japan). The surface wettability of the films was evaluated by water contact angle measurements.

3. Results and Discussion

3.1. Chemical Compositions and Phase Structures

Table 1 shows the co-sputtering parameters and chemical compositions of the prepared ternary TiNbTa and quaternary TiNbTaZr films. These alloys, with mixing entropy ranging from 0.54 to 1.36 R, are considered low- and medium-entropy alloys. The ternary TiNbTa films included a near-equiatomic Ti38Nb30Ta32, a Ti-enriched Ti83Nb10Ta7, an Nb-enriched Ti10Nb80Ta10, and a Ta-enriched Ti12Nb4Ta84 film. Figure 1a displays the XRD patterns of the TiNbTa films deposited on Ti-6Al-4V substrates. The Ti83Nb10Ta7 film exhibited a mixed hcp and bcc phase, whereas the Ti38Nb30Ta32, Ti10Nb80Ta10, and Ti12Nb4Ta84 films exhibited a bcc structure. The valence electron concentration (VEC) is an indicator for forecasting the crystalline phases of solid solutions in high entropy alloys [24,25]. The VEC values were 4.17, 4.61, 4.90, and 4.88 for the Ti83Nb10Ta7, Ti38Nb30Ta32, Ti10Nb80Ta10, and Ti12Nb4Ta84 films, respectively. Guo et al. [24] reported that bcc phases were stable as the VEC of alloys was less than 6.87, whereas Yuan et al. [25] have reported that a single bcc or hcp phase was stable as VEC >4.18 and <4.09, respectively, and a mixed bcc and hcp structure formed as 4.09 ≤ VEC < 4.18. In this study, the VEC of Ti83Nb10Ta7 film was 4.17, and it formed a mixed bcc and hcp phase, whereas the other three ternary films were stable in a bcc phase. Figure 2a displays the cross-sectional TEM (XTEM) image of the Ti83Nb10Ta7 film, which exhibits a typical columnar structure. The selected area electron diffraction (SAED) pattern indicates the coexistence of bcc and hcp phases (Figure 2b). Figure 2c shows the high-resolution TEM (HRTEM) image around the interface between the Ti83Nb10Ta7 film and C protective layer. As shown in the figure, the Ti83Nb10Ta7 film is crystalline, and the selected regions exhibit lattice fringes with d-spacing values of 0.234–0.235 nm in respect of bcc (110) planes. The lattice constants of the bcc phase in Ti83Nb10Ta7, Ti38Nb30Ta32, Ti10Nb80Ta10, and Ti12Nb4Ta84 films were determined to be 0.3303, 0.3289, 0.3300, and 0.3310 nm, respectively. The near-equiatomic Ti38Nb30Ta32 film possessed low lattice constants, whereas the Ta-enriched Ti12Nb4Ta84 film exhibited large lattice constants. On the other hand, the quaternary TiNbTaZr films were categorized as a near-equiatomic Ti33Nb19Ta21Zr27 film, a Ti-enriched Ti61Nb13Ta11Zr15 film, an Nb-enriched Ti15Nb68Ta8Zr9 film, a Zr-enriched Ti19Nb10Ta9Zr62 film, and a Ta-enriched Ti8Nb8Ta79Zr5 film. The low VEC values of 4.24 and 4.19 for the Ti61Nb13Ta11Zr15 and Ti19Nb10Ta9Zr62 films, respectively, imply the formation of a mixed bcc and hcp structure, which agrees with their XRD patterns (Figure 1b). In contrast, the Ti33Nb19Ta21Zr27, Ti15Nb68Ta8Zr9, and Ti8Nb8Ta79Zr5 films exhibited VEC values of 4.40, 4.76, and 4.87, respectively, and these films displayed a single bcc phase. The lattice constants of the bcc phase in Ti33Nb19Ta21Zr27, Ti15Nb68Ta8Zr9, and Ti8Nb8Ta79Zr5 films were determined to be 0.3368, 0.3322, and 0.3332 nm, respectively. If Vegard’s law is obeyed [23], the lattice constants of a solid solution can be calculated using the rule of mixture. The lattice constants of bcc Ti, Nb, Ta, and Zr are 0.332 [26,27], 0.33066 (International Center for Diffraction Data (ICDD) 00-035-0789), 0.33058 (ICDD 00-004-0788), and 0.35453 (ICDD 00-034-0657) nm, respectively. Figure 3 displays the relationship between the measured and calculated lattice constants of the bcc phases of the TiNbTa(Zr) films with a single bcc phase only, which shows a linear variation tendency. The fitted line slope of the measured to the calculated values was 0.9985.

3.2. Mechanical Properties

Table 2 lists the hardness (H) and Young’s modulus (E) of the Ti-6Al-4V substrate and TiNbTa(Zr) films. The H and E of the Ti-6Al-4V coupon were 4.9 and 124 GPa, respectively. A Young’s modulus of 110 GPa was reported for Ti-6Al-4V alloy [28]. TiNbSn [28] and TiNbTaZr [6] alloys with low Young’s modulus values of 40 and 60 GPa, respectively, were developed as alternative biocompatible materials. However, with the mechanisms of grain boundary strengthening and solid solution strengthening, metallic alloy films exhibit mechanical properties superior to those of bulk alloys [29]. The grain sizes evaluated from the FWHMs of (110) reflections in Bragg–Brentano XRD patterns were 21, 37, 24, and 32 nm for the Ti83Nb10Ta7, Ti38Nb30Ta32, Ti10Nb80Ta10, and Ti12Nb4Ta84 films, respectively, and 22, 33, 34, 33, and 38 nm for the Ti61Nb13Ta11Zr15, Ti61Nb13Ta11Zr15, Ti15Nb68Ta8Zr9, Ti19Nb10Ta9Zr62, and Ti8Nb8Ta79Zr5 films, respectively. Because these films exhibited various compositions, the relationship between hardness and grain size could not be interpreted by the normal or inverse Hall–Petch strengthening (grain boundary strengthening) mechanism. The H values were 4.0, 4.3, 11.7, and 6.9 GPa for hcp-Ti, bcc-Nb, bcc-Ta, and hcp-Zr films, respectively. The films with high Ta content should have high H values. Ti12Nb4Ta84 and Ti8Nb8Ta79Zr5 films exhibited maximum hardness values of 7.7 and 12.1 GPa among the ternary and quaternary films, respectively, which implies that the solid solution strengthening mechanism dominated the hardness of these TiNbTa(Zr) films. As shown in Table 2, the TiNbTa(Zr) films with a single bcc phase exhibited high E values of 142–226 GPa, whereas the films with a mixture of bcc and hcp phases exhibited low E values of 108–115 GPa. This variation was not consistent with that reported for bulk Ti alloys, i.e., β(bcc)-phase Ti-alloys that exhibited low Young’s modulus. Though these TiNbTa(Zr) films exhibited high E values, these films should not cause a stress-shielding effect due to the low volume ratios of film-to-bulk Ti-6Al-4V. Moreover, selected TiNbTa(Zr) films improved the toughness and resistance to plastic deformation represented by high H/E [30] and H3/E2 [31] indicators, respectively. The Ti12Nb4Ta84 and Ti8Nb8Ta79Zr5 films exhibited favorable H/E and H3/E2 values in the studied ternary and quaternary films, respectively.

3.3. Corrosion Resistance

Figure 4 depicts the potentiodynamic polarization curves of TiNbTa and TiNbTaZr films and the uncoated Ti-6Al-4V substrate as immersed in a 3.5 wt.% NaCl aqueous solution. The corrosion potential (Ecorr) and corrosion current density (Icorr) values, determined by the Tafel extrapolation method, are presented in Table 3. The polarization resistance (Rp) values of the samples were determined according to the Stern–Geary equation [32]. The Rp values of TiNbTa films ranged from 4.1 × 103 to 1.5 × 104 KΩcm2, whereas the Rp values of TiNbTaZr films were in the range of 5.3 × 103–2.9 × 104 KΩcm2. Although there are no significant pitting behaviors for Ti83Nb10Ta7 and Ti61Nb13Ta11Zr15, it is noteworthy that the ternary and quaternary Ti-enriched films have the Rp values of 4.1 and 5.3 × 103 KΩcm2, respectively. These Rp values are similar to the bare Ti-6Al-4V of 4.5 × 103 KΩcm2, which confirmed that the Ti-enriched films had inferior corrosion resistance in NaCl solution in relation to the other TiNbTa(Zr) films. The enhancement of Rp values caused by adding Zr into ternary TiNbTa films is evident. For example, the Rp ratio of Ti-enriched Ti83Nb10Ta7 related to the bare Ti-6Al-4V was 0.9, whereas that of the Ti-enriched Ti61Nb13Ta11Zr15 was 1.2. Moreover, the Rp ratio increased from 1.7 for the Nb-enriched Ti10Nb80Ta10 to 4.3 for the Nb-enriched Ti15Nb68Ta8Zr9. Additionally, the Rp ratio increased from 3.3 for the Ta-enriched Ti12Nb4Ta84 to 4.3 for the Ta-enriched Ti8Nb8Ta79Zr5. The change in corrosion resistance was associated with the fact that ZrO2 has the more negative Gibbs free energy per mole of O2 among the available four oxides ZrO2, TiO2, Ta2O5, and Nb2O5, implying that the passivation film becomes thicker with an increase in Zr content [33]. Therefore, the Icorr of quaternary films was lower than that of ternary films, resulting in higher corrosion resistance.
The Ti33Nb19Ta21Zr27 film and the bare Ti-6Al-4V exhibited Ecorr values of −281 and −250 mV, respectively. The Ti33Nb19Ta21Zr27 film exhibited a more negative corrosion potential value than that of bare Ti-6Al-4V and had the highest Rp of 2.9 × 104 KΩcm2 or Rp ratio of 6.4 in this study. The increasing corrosion resistance of Ti33Nb19Ta21Zr27 can be attributed to the formation of amorphous films comprising Ta2O5, Nb2O5, and ZrO2 on the free surface. These aforementioned dense and stable passive oxide layers can restrict the electrochemical reactions, leading to a decrease in corrosion current density [18]. However, the Ti19Nb10Ta9Zr62 film with the highest Zr content in this study exhibited current fluctuations in its potentiodynamic polarization curve similar to that observed for the uncoated Ti-6Al-4V (Figure 4), implying the conduct of pitting behavior. According to the previous literature [34,35], the possible explanation for the pitting behavior can be ascribed to the presence of chloride ions in NaCl solution. The aggressive Cl ions with high electronegativity can easily combine with oxygen vacancies in the passive layer, which results in the formation of localized pitting corrosion at the grain boundaries and induces the formation of a porous structure. Moreover, the galvanic corrosion that occurs at the interphase of bcc and hcp phase boundaries in Ti-6Al-4V alloy can accelerate the dissolving rate of the passivation film [33]. The results of the corrosion test in NaCl solution are as follows: (1) the quaternary TiNbTaZr films presented higher corrosion resistance than that of the ternary TiNbTa films, (2) the Ti-enriched Ti61Nb13Ta11Zr15 film revealed a corrosion resistance slightly higher than that of the bare Ti-6Al-4V, and (3) the Zr-enriched Ti19Nb10Ta9Zr62 film revealed pitting behavior. Therefore, the Ti-6Al-4V modified with Ti33Nb19Ta21Zr27, Ti15Nb68Ta8Zr9, and Ti8Nb8Ta79Zr5 films should be suitable assemblies for biocompatible implantation in this study, and their high oxidation resistance was attributed to the formation of the stable passivation film on the single-bcc-phase film.

3.4. Elution Test and Formation of Surficial Oxide Layers

Figure 5 shows the images of water contact angles between the water drop and the TiNbTa(Zr) films and bare Ti-6Al-4V substrate. The contact angle was 50° for the Ti-6Al-4V substrate, which was comparable to the 55° contact angle reported in a previous study [36]. With a hydrophilic surface, the Ti-6Al-4V alloy enhanced the adsorption of human osteoblast cells and subsequent cell growth. The contact angles were 85° for the Ti83Nb10Ta7, Ti38Nb30Ta32, and Ti10Nb80Ta10 films, whereas the Ti12Nb4Ta84 film exhibited a lower contact angle of 77°. The passive oxide films formed on the metallic surfaces affected the water contact angle values by changing the surface energies [37]. Though the contact angles increased after the Ti-6Al-4V alloy was covered with TiNbTa films, these samples remained hydrophilic. The contact angles of Ti61Nb13Ta11Zr15, Ti33Nb19Ta21Zr27, Ti15Nb68Ta8Zr9, Ti19Nb10Ta9Zr62, and Ti8Nb8Ta79Zr5 film were 70°, 61°, 75°, 72°, and 70°, respectively, which were lower than those of the TiNbTa films.
Figure 6a depicts the XTEM image of the Ti33Nb19Ta21Zr27 film after immersion in Ringer’s solution for eight weeks. The Ti33Nb19Ta21Zr27 film exhibited a typical columnar structure and the SAED pattern exhibited a bcc structure (Figure 6b). Figure 6c displays the HRTEM image at the near-surface region. An amorphous oxide layer of approximately 10 nm thickness was observed at the free surface. No evident lattice fringes were observed in the amorphous oxide layer. Crystalline domains with d-spacing values of 0.234–0.235 nm were observed beneath the oxide layer, which represented the bcc (110) planes.
Figure 7 shows the compositional profiles of the TiNbTa and TiNbTaZr films examined by XPS after these samples were immersed in Ringer’s solutions for eight weeks. The near-equiatomic Ti38Nb30Ta32 and Ti33Nb19Ta21Zr27 films exhibited inward diffusion of O and Al during the elution test (Figure 7a,b). The O contents of the Ti38Nb30Ta32 and Ti33Nb19Ta21Zr27 films exhibited a common decreasing trend with an increase in the film depth. The presence of Al at the near-surface regions of 0–7.8 nm and 0–11.7 nm for the Ti38Nb30Ta32 and Ti33Nb19Ta21Zr27 films, respectively, resulted from the inward diffusion of Al ions from the Ringer’s solution. Beneath the near-surface regions, no Al or V atoms were observed, which implies that the Ti38Nb30Ta32 and Ti33Nb19Ta21Zr27 films formed protective barriers to isolate the Ti-6Al-4V substrate. Figure 7c displays the compositional profiles of the Ti8Nb8Ta79Zr5 film. The inward diffusion of O was restricted to a shallower depth of the Ti8Nb8Ta79Zr5 film related to those of the near-equiatomic Ti38Nb30Ta32 and Ti33Nb19Ta21Zr27 films, which implies a higher oxidation resistance due to the formation of Ta2O5. In contrast, O diffused into the Ti-6Al-4V substrate (Figure 7d), which was attributed to the lack of a passivation oxide film such as Ta2O5 and Nb2O5 [3].
Figure 8 depicts the XPS profiles of the Ti38Nb30Ta32 film at depths of 0, 7.8, and 47.7 nm after immersion in Ringer’s solution for eight weeks. The splitting energies were 5.54, 2.72, and 1.91 eV for Ti 2p, Nb 3d, and Ta 4f doublets [38], respectively. The area ratios were set as 2:1, 3:2, and 4:3 for 2p3/2:2p1/2, 3d5/2:3d3/2, and 4f7/2:4f5/2, respectively. On the free surface, binding energies of 456.55, 205.23, 24.27, and 72.50 eV indicated the signals of Ti 2p3/2 in a TiO2 form, Nb 3d5/2 in a Nb2O5 form, Ta 4f7/2 in a TaO2 form, and Al 1s in a Al2O3 form, respectively. The Ti4+ 2p3/2 signal exhibited a shift of −2.25 eV related to 458.8 eV shown in the handbook [38]. The FWHM values of Ti4+ 2p3/2 and 2p1/2 were 1.31 and 2.20 eV, respectively. A wider FWHM value for Ti 2p1/2 than that for the Ti 2p3/2 that has been reported in the literature [39,40]. The aforementioned FWHM ratio was sustained for analyzing the other Ti doublets in this study. In contrast, both the FWHM ratios of 3d5/2:3d3/2 and 4f7/2:4f5/2 for Nb and Ta doublets, respectively, were set as 1:1. At a depth of 7.8 nm beneath the free surface, the Ti signal consisted of Ti4+, Ti3+, and Ti2+ doublets, the Nb signal consisted of Nb5+, Nb4+, Nb2+, and Nb0 doublets, and the Ta signal consisted of Ta5+, Ta4+, Ta2+, and Ta0 doublets, whereas the Al signal was not obvious. At a depth of 47.7 nm, only Ti2+, Ti0, Nb2+, Nb0, Ta2+, and Ta0 doublets were observed, and no Al signal was detectable. The decrease in the intensity of Al signals along the depth direction of the Ti38Nb30Ta32 film suggested that the Al ions were contributed from the Ringer’s solution during the elution test as examined from the compositional profiles (Figure 7). The decrease in O content in the depth direction of the Ti38Nb30Ta32 film (Figure 7a) resulted in the elements of the Ti38Nb30Ta32 film varying from full oxidization at the free surface to a mixture with various oxidization states and to the state being dominated with metallic atoms accompanied by minor amounts of divalent ions. Figure 9 exhibits the XPS depth profiles of the Ti-6Al-4V substrate after immersion in Ringer’s solution for eight weeks. The shift of Ti4+ 2p3/2 signal was −2.61 eV in relation to 458.8 eV. A Ti4+ doublet was observed at the free surface and depths of 7.8 and 47.7 nm, whereas a minor Ti3+ doublet was examined at a depth of 47.7 nm. Both the Al bondings in Al2O3 and Al-halides [38] were detected. These Al atoms were eluted from the bare Ti-6Al-4V substrate without protective films.

4. Conclusions

The assemblies of TiNbTa(Zr) films/Ti-6Al-4V alloys utilized as biocompatible implant materials were evaluated in this study. The chemical compositions of the TiNbTa(Zr) films were regulated through co-sputtering with four sputter sources. The phase constitutions of TiNbTa(Zr) films were correlated to their VEC values. The TiNbTa(Zr) films with a VEC value of 4.17–4.24 exhibited a mixture of hcp and bcc phases, whereas the films with a VEC value of 4.40–4.90 revealed a single bcc phase. The solid solution strengthening mechanism dominated the hardness of these TiNbTa(Zr) films. In contrast to the performance of bulk Ti alloys, β-phase TiNbTa(Zr) films exhibited Young’s modulus values of 142–226 GPa, which were higher than those (108–115 GPa) of the films with an α–β mixed phase. The corrosion test in a 3.5 wt.% NaCl aqueous solution indicated that the Ti33Nb19Ta21Zr27, Ti15Nb68Ta8Zr9, and Ti8Nb8Ta79Zr5 films exhibited high anticorrosive properties deriving from the stable passivation film comprising constitutions of Ta2O5, Nb2O5, and ZrO2. The water contact angles of TiNbTaZr films were lower than those of the TiNbTa films, which implies that the TiNbTaZr films were more hydrophilic. The XPS depth profiles indicated that the elution of Al atoms from the Ti-6Al-4V alloy was inhibited by the TiNbTa(Zr) films as the TiNbTa(Zr)/Ti-6Al-4V assemblies were immersed in Ringer’s solution at 310 K for eight weeks, which was attributed to the formation of an amorphous surficial oxide layer consisting of Ta2O5 and Nb2O5. In summary, the Ti33Nb19Ta21Zr27, Ti15Nb68Ta8Zr9, and Ti8Nb8Ta79Zr5 films are suitable candidates for protective coatings on Ti-6Al-4V alloys when applied as biocompatible materials.

Author Contributions

Conceptualization, Y.-J.C. and Y.-I.C.; funding acquisition, Y.-I.C. and L.-C.C.; investigation, C.-Y.L. and Y.-J.C.; resources, Y.-I.C. and L.-C.C.; validation, C.-Y.L.; visualization, C.-Y.L.; writing—original draft preparation, C.-Y.L.; writing—review and editing, Y.-I.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science and Technology, Taiwan, grant numbers 110-2221-E-019-015 and 110-2221-E-131-013. The APC was funded by National Taiwan Ocean University.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors thank the Instrumentation Center at the National Tsing Hua University for EPMA and XPS analyses and the Joint Center for High Valued Instruments at NSYSU for the FIB operation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rodil, S.E.; Olivares, R.; Arzate, H.; Muhl, S. Properties of carbon films and their biocompatibility using in-vitro tests. Diam. Relat. Mater. 2003, 12, 931–937. [Google Scholar] [CrossRef]
  2. Sharma, A.; Oh, M.C.; Kim, J.-T.; Srivastava, A.K.; Ahn, B. Investigation of electrochemical corrosion behavior of additive manufactured Ti–6Al–4V alloy for medical implants in different electrolytes. J. Alloys Compd. 2020, 830, 154620. [Google Scholar] [CrossRef]
  3. Hussein, A.H.; Gepreel, M.A.; Gouda, M.K.; Hefnawy, A.M.; Kandil, S.H. Biocompatibility of new Ti–Nb–Ta base alloys. Mater. Sci. Eng. C 2016, 61, 574–578. [Google Scholar] [CrossRef] [PubMed]
  4. Williams, D.F. On the mechanisms of biocompatibility. Biomaterials 2008, 29, 2941–2953. [Google Scholar] [CrossRef]
  5. Van Lenthe, G.H.; de Waal Malefijt, M.C.; Huiskes, R. Stress shielding after total knee replacement may cause bone resorption in the distal femur. J. Bone Joint Surg. 1997, 79, 117–122. [Google Scholar] [CrossRef]
  6. Niinomi, M.; Nakai, M. Titanium-based biomaterials for preventing stress shielding between implant devices and bone. Int. J. Biomater. 2011, 2011, 836587. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Huiskes, R.; Weinans, H.; Rietbergen, B.V. The relationship between stress shielding and bone resorption around total hip stems and the effects of flexible materials. Clin. Orthop. Relat. Res. 1992, 274, 124–134. [Google Scholar] [CrossRef] [Green Version]
  8. Glassman, A.H.; Bobyn, J.D.; Tanzer, M. New femoral designs: Do they influence stress shielding? Clin. Orthop. Relat. Res. 2006, 453, 64–74. [Google Scholar] [CrossRef] [PubMed]
  9. Nagels, J.; Stokdijk, M.; Rozing, P.M. Stress shielding and bone resorption in shoulder arthroplasty. J. Shoulder Elbow Surg. 2003, 12, 35–39. [Google Scholar] [CrossRef]
  10. Vilhena, L.M.; Shumayal, A.; Ramalho, A.; Ferreira, J.A.M. Tribocorrosion behaviour of Ti6Al4V produced by selective laser melting for dental implants. Lubricants 2020, 8, 22. [Google Scholar] [CrossRef]
  11. Fellah, M.; Labaïz, M.; Assala, O.; Dekhil, L.; Taleb, A.; Rezag, H.; Iost, A. Tribological behavior of Ti-6Al-4V and Ti-6Al-7Nb alloys for total hip prosthesis. Adv. Tribol. 2014, 2014, 451387. [Google Scholar] [CrossRef] [Green Version]
  12. Kuroda, D.; Niinomi, M.; Morinaga, M.; Kato, Y.; Yashiro, T. Design and mechanical properties of new β type titanium alloys for implant materials. Mater. Sci. Eng. A 1998, 243, 244–249. [Google Scholar] [CrossRef]
  13. Zaffe, D.; Bertoldi, C.; Consolo, U. Accumulation of aluminum in lamellar bone after implantation of titanium plates, Ti–6Al–4V screws, hydroxyapatite granules. Biomaterials 2004, 25, 3837–3844. [Google Scholar] [CrossRef] [PubMed]
  14. Raabe, D.; Sander, B.; Friák, M.; Ma, D.; Neugebauer, J. Theory-guided bottom-up design of β-titanium alloys as biomaterials based on first principles calculations: Theory and experiments. Acta Mater. 2007, 55, 4475–4487. [Google Scholar] [CrossRef]
  15. Song, Y.; Xu, D.S.; Yang, R.; Li, D.; Wu, W.T.; Guo, Z.X. Theoretical study of the effects of alloying elements on the strength and modulus of β-type bio-titanium alloys. Mater. Sci. Eng. A 1999, 260, 269–274. [Google Scholar] [CrossRef]
  16. Niinomi, M.; Kuroda, D.; Fukunaga, K.; Morinaga, M.; Kato, Y.; Yashiro, T.; Suzuki, A. Corrosion wear fracture of new β type biomedical titanium alloys. Mater. Sci. Eng. A 1999, 263, 193–199. [Google Scholar] [CrossRef]
  17. Biesiekierski, A.; Wang, J.; Gepreel, M.A.; Wen, C. A new look at biomedical Ti-based shape memory alloys. Acta Biomater. 2012, 8, 1661–1669. [Google Scholar] [CrossRef]
  18. Chen, Y.H.; Chuang, W.S.; Huang, J.C.; Wang, X.; Chou, H.S.; Lai, Y.J.; Lin, P.H. On the bio-corrosion and biocompatibility of TiTaNb medium entropy alloy films. Appl. Surf. Sci. 2020, 508, 145307. [Google Scholar] [CrossRef]
  19. Lai, B.-W.; Chang, Y.-Y.; Shieh, T.-M.; Huang, H.-L. Biocompatibility and microstructure-based stress analyses of TiNbZrTa composite films. Materials 2022, 15, 29. [Google Scholar] [CrossRef]
  20. Tüten, N.; Canadinc, D.; Motallebzadeh, A.; Bal, B. Microstructure and tribological properties of TiTaHfNbZr high entropy alloy coatings deposited on Ti–6Al–4V substrates. Intermetallics 2019, 105, 99–106. [Google Scholar] [CrossRef]
  21. Peighambardoust, N.S.; Alamdari, A.A.; Unal, U.; Motallebzadeh, A. In vitro biocompatibility evaluation of Ti1.5ZrTa0.5Nb0.5Hf0.5 refractory high-entropy alloy film for orthopedic implants: Microstructural, mechanical properties and corrosion behavior. J. Alloys Compd. 2021, 883, 160786. [Google Scholar] [CrossRef]
  22. Chen, Y.I.; Chen, C.Y.; Chang, L.C.; Kai, W. Characterization of cosputtered NbTaMoW films. J. Mater. Res. Technol. 2021, 15, 1090–1099. [Google Scholar] [CrossRef]
  23. Cullity, B.D.; Stock, S.R. Elements of X-ray Diffraction, 3rd ed.; Prentice-Hall: Hoboken, NJ, USA, 2001. [Google Scholar]
  24. Guo, S.; Ng, C.; Lu, J.; Liu, C.T. Effect of valence electron concentration on stability of fcc or bcc phase in high entropy alloys. J. Appl. Phys. 2011, 109, 103505. [Google Scholar] [CrossRef] [Green Version]
  25. Yuan, Y.; Wu, Y.; Yang, Z.; Liang, X.; Lei, Z.; Huang, H.; Wang, H.; Liu, X.; An, K.; Wu, W.; et al. Formation, structure and properties of biocompatible TiZrHfNbTa high-entropy alloys. Mater. Res. Lett. 2019, 7, 225–231. [Google Scholar] [CrossRef] [Green Version]
  26. Han, G.; Lu, X.; Xia, Q.; Lei, B.; Yan, Y.; Shang, C.J. Face-centered-cubic titanium - A new crystal structure of Ti in a Ti- 8Mo-6Fe alloy. J. Alloys Compd. 2018, 748, 943–952. [Google Scholar] [CrossRef]
  27. Hua, K.; Zhang, Y.; Kou, H.; Li, J.; Gan, W.; Fundenberger, J.J.; Esling, C. Composite structure of α phase in metastable β Ti alloys induced by lattice strain during β to α phase transformation. Acta Mater. 2017, 132, 307–326. [Google Scholar] [CrossRef]
  28. Miura, K.; Yamada, N.; Hanada, S.; Jung, T.-K.; Itoi, E. The bone tissue compatibility of a new Ti–Nb–Sn alloy with a low Young’s modulus. Acta Biomater. 2011, 7, 2320–2326. [Google Scholar] [CrossRef]
  29. Song, B.; Li, Y.; Cong, Z.; Li, Y.; Song, Z.; Chen, J. Effects of deposition temperature on the nanomechanical properties of refractory high entropy TaNbHfZr films. J. Alloys Compd. 2019, 797, 1025–1030. [Google Scholar] [CrossRef]
  30. Pogrebnjak, A.D.; Beresnev, V.M.; Bondar, O.V.; Postolnyi, B.O.; Zaleski, K.; Coy, E.; Jurga, S.; Lisovenko, M.O.; Konarski, P.; Rebouta, L.; et al. Superhard CrN/MoN coatings with multilayer architecture. Mater. Des. 2018, 153, 47–59. [Google Scholar] [CrossRef]
  31. Musil, J. Hard nanocomposite coatings: Thermal stability, oxidation resistance and toughness. Surf. Coat. Technol. 2012, 207, 50–65. [Google Scholar] [CrossRef]
  32. Stern, M. A method for determining corrosion rates from linear polarization data. Corrosion 1958, 14, 60–64. [Google Scholar] [CrossRef]
  33. Ji, P.F.; Li, B.; Chen, B.H.; Wang, F.; Ma, W.; Zhang, X.Y.; Ma, M.Z.; Liu, R.P. Effect of Nb addition on the stability and biological corrosion resistance of Ti-Zr alloy passivation films. Corros. Sci. 2020, 170, 108696. [Google Scholar] [CrossRef]
  34. Alves, V.A.; Reis, R.Q.; Santos, I.C.B.; Souza, D.G.; de F. Gonçalves, T.; Pereira-da-Silva, M.A.; Rossi, A.; da Silva, L.A. In situ impedance spectroscopy study of the electrochemical corrosion of Ti and Ti–6Al–4V in simulated body fluid at 25 °C and 37 °C. Corros. Sci. 2009, 51, 2473–2482. [Google Scholar] [CrossRef]
  35. Jiang, Z.; Dai, X.; Norby, T.; Middleton, H. Investigation of pitting resistance of titanium based on a modified point defect model. Corros. Sci. 2011, 53, 815–821. [Google Scholar] [CrossRef]
  36. Mirhosseini, N.; Crouse, P.L.; Schmidth, M.J.J.; Li, L.; Garrod, D. Laser surface micro-texturing of Ti–6Al–4V substrates for improved cell integration. Appl. Surf. Sci. 2007, 253, 7738–7743. [Google Scholar] [CrossRef]
  37. Razazzadeh, A.; Atapour, M.; Enayati, M.H. Corrosion characteristics of TiNbMoMnFe high entropy thin film deposited on AISI316L for biomedical applications. Met. Mater. Int. 2021, 27, 2341–2352. [Google Scholar] [CrossRef]
  38. Moulder, J.F.; Stickle, W.F.; Sobol, P.E.; Bomben, K.D. Handbook of X-ray Photoelectron Spectroscopy; Chastain, J., King, R.C., Eds.; Physical Electronics: Chanhassen, MN, USA, 1995. [Google Scholar]
  39. Krishna, D.N.G.; George, R.P.; Philip, J. Determination of nanoscale titanium oxide thin film phase composition using X-ray photoelectron spectroscopy valence band analysis. Thin Solid Films 2019, 681, 58–68. [Google Scholar] [CrossRef]
  40. Chi, M.; Sun, X.; Lozano-Blanco, G.; Tatarchuk, B.J. XPS and FTIR investigations of the transient photocatalytic decomposition of surface carbon contaminants from anatase TiO2 in UHV starved water/oxygen environments. Appl. Surf. Sci. 2021, 570, 151147. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of (a) TiNbTa and (b) TiNbTaZr films.
Figure 1. XRD patterns of (a) TiNbTa and (b) TiNbTaZr films.
Coatings 12 01985 g001
Figure 2. (a) XTEM image, (b) SAED pattern, and (c) HRTEM image of the Ti83Nb10Ta7 film prepared on Si substrate.
Figure 2. (a) XTEM image, (b) SAED pattern, and (c) HRTEM image of the Ti83Nb10Ta7 film prepared on Si substrate.
Coatings 12 01985 g002aCoatings 12 01985 g002b
Figure 3. Relationship between the measured and calculated lattice constants of the bcc phases of the TiNbTa(Zr) films.
Figure 3. Relationship between the measured and calculated lattice constants of the bcc phases of the TiNbTa(Zr) films.
Coatings 12 01985 g003
Figure 4. Potentiodynamic polarization curves of (a) TiNbTa and (b) TiNbTaZr films and the Ti-6Al-4V substrate in a 3.5 wt.% NaCl aqueous solution.
Figure 4. Potentiodynamic polarization curves of (a) TiNbTa and (b) TiNbTaZr films and the Ti-6Al-4V substrate in a 3.5 wt.% NaCl aqueous solution.
Coatings 12 01985 g004
Figure 5. Water contact angles on (a) Ti38Nb30Ta32, (b) Ti12Nb4Ta84, (c) Ti33Nb19Ta21Zr27, and (d) Ti8Nb8Ta79Zr5 films and (e) bare Ti–6Al–4V substrate.
Figure 5. Water contact angles on (a) Ti38Nb30Ta32, (b) Ti12Nb4Ta84, (c) Ti33Nb19Ta21Zr27, and (d) Ti8Nb8Ta79Zr5 films and (e) bare Ti–6Al–4V substrate.
Coatings 12 01985 g005
Figure 6. (a) XTEM image, (b) SAED pattern, and (c) HRTEM image of the Ti33Nb19Ta21Zr27 film after immersion in Ringer’s solution for eight weeks.
Figure 6. (a) XTEM image, (b) SAED pattern, and (c) HRTEM image of the Ti33Nb19Ta21Zr27 film after immersion in Ringer’s solution for eight weeks.
Coatings 12 01985 g006
Figure 7. Compositional profiles of (a) Ti38Nb30Ta32, (b) Ti33Nb19Ta21Zr27, and (c) Ti8Nb8Ta79Zr5 films and (d) Ti-6Al-4V substrate examined by XPS after immersion in Ringer’s solution for eight weeks.
Figure 7. Compositional profiles of (a) Ti38Nb30Ta32, (b) Ti33Nb19Ta21Zr27, and (c) Ti8Nb8Ta79Zr5 films and (d) Ti-6Al-4V substrate examined by XPS after immersion in Ringer’s solution for eight weeks.
Coatings 12 01985 g007
Figure 8. XPS analysis results of (a) Ti, (b) Nb, (c) Ta, and (d) Al profiles at depths of 0, 7.8, and 47.7 nm of the Ti38Nb30Ta32 film after immersion in Ringer’s solution for eight weeks.
Figure 8. XPS analysis results of (a) Ti, (b) Nb, (c) Ta, and (d) Al profiles at depths of 0, 7.8, and 47.7 nm of the Ti38Nb30Ta32 film after immersion in Ringer’s solution for eight weeks.
Coatings 12 01985 g008
Figure 9. XPS analysis results of (a) Ti and (b) Al profiles at depths of 0, 7.8, and 47.7 nm of the Ti-6Al-4V substrate after immersion in Ringer’s solution for eight weeks.
Figure 9. XPS analysis results of (a) Ti and (b) Al profiles at depths of 0, 7.8, and 47.7 nm of the Ti-6Al-4V substrate after immersion in Ringer’s solution for eight weeks.
Coatings 12 01985 g009
Table 1. Co-sputtering parameters and chemical compositions of TiNbTa and TiNbTaZr films.
Table 1. Co-sputtering parameters and chemical compositions of TiNbTa and TiNbTaZr films.
SampleSputter Power (W)Chemical Composition (at.%)ΔSmix 1
PTiPNbPTaPZrTiNbTaZrO(R)
Ti83Nb10Ta7200201076.3 ± 0.59.3 ± 0.16.9 ± 0.07.5 ± 0.40.58
Ti38Nb30Ta32180807036.4 ± 0.128.5 ± 0.130.1 ± 0.15.0 ± 0.11.09
Ti10Nb80Ta1030200209.2 ± 0.176.2 ± 0.310.0 ± 0.04.6 ± 0.40.64
Ti12Nb4Ta84702020010.9 ± 0.24.2 ± 0.678.2 ± 1.36.7 ± 2.00.54
Ti61Nb13Ta11Zr1520030203061.3 ± 0.512.6 ± 0.211.3 ± 0.314.8 ± 0.01.09
Ti33Nb19Ta21Zr27180807014032.8 ± 0.218.9 ± 0.121.0 ± 0.327.3 ± 0.11.36
Ti15Nb68Ta8Zr970200203014.5 ± 0.368.2 ± 0.38.1 ± 0.49.2 ± 0.20.96
Ti19Nb10Ta9Zr6270302003019.1 ± 0.210.5 ± 0.18.5 ± 0.061.9 ± 0.11.06
Ti8Nb8Ta79Zr57030202007.6 ± 0.28.1 ± 0.178.9 ± 0.65.4 ± 0.60.74
1 ΔSmix: Mixing entropy.
Table 2. Mechanical properties of TiNbTa and TiNbTaZr films.
Table 2. Mechanical properties of TiNbTa and TiNbTaZr films.
SampleH 1 (GPa)E 2 (GPa)H/EH3/E2 (GPa)
Ti-6Al-4V4.9 ± 0.4124 ± 80.0400.008
Ti83Nb10Ta72.7 ± 0.3110 ± 70.0250.002
Ti38Nb30Ta325.0 ± 0.2150 ± 100.0330.006
Ti10Nb80Ta104.2 ± 0.5142 ± 80.0300.004
Ti12Nb4Ta847.7 ± 0.3216 ± 60.0360.010
Ti61Nb13Ta11Zr153.9 ± 1.0108 ± 170.0360.005
Ti33Nb19Ta21Zr275.9 ± 0.9163 ± 120.0360.008
Ti15Nb68Ta8Zr99.6 ± 1.1165 ± 140.0580.032
Ti19Nb10Ta9Zr627.0 ± 0.3115 ± 40.0610.026
Ti8Nb8Ta79Zr512.1 ± 1.0226 ± 90.0540.035
1 H: Hardness; 2 E: Young’s modulus.
Table 3. Corrosion characteristics of the Ti-6Al-4V substrate and TiNbTa(Zr) films.
Table 3. Corrosion characteristics of the Ti-6Al-4V substrate and TiNbTa(Zr) films.
SampleEcorr (mV)Icorr (μA/cm2)Rp (KΩcm2)Rp Ratio
Ti-6Al-4V−2500.0044.5 × 1031.0
Ti83Nb10Ta7−2680.0064.1 × 1030.9
Ti38Nb30Ta32−3150.0028.8 × 1032.0
Ti10Nb80Ta10−2140.0037.6 × 1031.7
Ti12Nb4Ta84−3750.0011.5 × 1043.3
Ti61Nb13Ta11Zr15−3080.0055.3 × 1031.2
Ti33Nb19Ta21Zr27−2810.0012.9 × 1046.4
Ti15Nb68Ta8Zr9−3650.0012.1 × 1044.7
Ti19Nb10Ta9Zr62−2470.0027.7 × 1031.7
Ti8Nb8Ta79Zr5−4410.0011.9 × 1044.2
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chen, Y.-I.; Chen, Y.-J.; Lai, C.-Y.; Chang, L.-C. Mechanical and Anticorrosive Properties of TiNbTa and TiNbTaZr Films on Ti-6Al-4V Alloy. Coatings 2022, 12, 1985. https://doi.org/10.3390/coatings12121985

AMA Style

Chen Y-I, Chen Y-J, Lai C-Y, Chang L-C. Mechanical and Anticorrosive Properties of TiNbTa and TiNbTaZr Films on Ti-6Al-4V Alloy. Coatings. 2022; 12(12):1985. https://doi.org/10.3390/coatings12121985

Chicago/Turabian Style

Chen, Yung-I, Yi-Jyun Chen, Cheng-Yi Lai, and Li-Chun Chang. 2022. "Mechanical and Anticorrosive Properties of TiNbTa and TiNbTaZr Films on Ti-6Al-4V Alloy" Coatings 12, no. 12: 1985. https://doi.org/10.3390/coatings12121985

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