Next Article in Journal
Porous Glass with Layered Morphology Prepared by Phase Separation
Previous Article in Journal
Effortless Fabrication of Nanofused HKUST-1 for Enhanced Catalytic Efficiency in the Cyanosilylation of Aldehyd
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Annealing Time on Corrosion Behaviours of Zr56Cu19Ni11Al9Nb5 in Hank Solution

1
School of Materials Science and Engineering, Wuhan University of Technology, Wuhan 430070, China
2
School of Energy and Power Engineering, Huazhong University of Science and Technology, Wuhan 430074, China
3
Hubei Hongle Cable Co., Ltd., Honghu, Jingzhou 433225, China
*
Author to whom correspondence should be addressed.
Materials 2025, 18(5), 1132; https://doi.org/10.3390/ma18051132
Submission received: 23 November 2024 / Revised: 17 February 2025 / Accepted: 23 February 2025 / Published: 3 March 2025
(This article belongs to the Section Corrosion)

Abstract

:
The microstructures of the as-cast and annealed Zr56Cu19Ni11Al9Nb5 were investigated by X-ray diffraction (XRD) and scanning electron microscopy (SEM), their microhardness values were tested, and their corrosion behaviours in Hank solution were studied. XRD results and SEM analysis showed that the as-cast sample was amorphous, and crystallisation occurred in the samples annealed at 923 K for 5–30 min with crystals of Zr2Cu and Zr2Ni. Microhardness gradually increased and then levelled off, due to higher crystallisation degree with longer annealing time. Passivation occurred for all the samples in Hank solution. Prolonged annealing time leads to the initial rise and then a drop in corrosion resistance. Annealing for 5 min resulted in the highest corrosion resistance, with high corrosion potential Ecorr at −0.007 VSCE, versus saturated calomel electrode (SCE), i.e., 0.234 VSHE, versus standard hydrogen electrode (SHE), the smallest corrosion current density icorr at 2.20 × 10−7 A·cm−2, the highest pitting potential Epit at 0.415 VSCE (i.e., 0.656 VSHE), the largest passivation region Epit–Ecorr at 0.421 VSHE, the largest arc radius, and the largest sum of charge transfer resistance and film resistance Rct + Rf at 15489 Ω·cm2. Annealing for 30 min led to the lowest corrosion resistance, with low Ecorr at −0.069 VSCE (i.e., 0.172 VSHE), large icorr at 1.32 × 10−6 A·cm−2, low Epit at −0.001 VSCE (i.e., 0.240 VSHE), small Epit − Ecorr at 0.068 VSHE, the smallest arc radius, and the smallest Rct + Rf at 4070 Ω·cm2. When the annealing time was appropriate, the homogeneous microstructure of nanocrystals in an amorphous matrix resulted in improved passivation film, leading to the rise of corrosion resistance. However, if the annealing time was prolonged, the inhomogeneous microstructure of larger crystals in an amorphous matrix resulted in a drop in corrosion resistance. Localised corrosion was observed, with corrosion products of ZrO2, Cu2O, CuO, Ni(OH)2, Al2O3, and Nb2O5.

1. Introduction

Bulk metallic glasses (BMGs) and bulk metallic glass composites (BMGCs) have been widely studied [1,2,3]. Zr-based BMGs showed promising biomedical applications due to high strength and hardness, good corrosion resistance, good glass forming ability, good biocompatibility, good wear resistance, and low elastic modulus [4,5]. Zr61Cu17.5Ni10Al7.5Si4 exhibited good corrosion resistance in Hank solution, and good biocompatibility, indicating good potential for biomedical applications [6]. Corrosion resistance is affected by chemical composition. Proper addition of Nb and Ti led to better corrosion resistance [7,8]. Zr55Al20Co25−xNbx (x = 0, 2.5, 5) displayed good corrosion resistance in Hank solution and phosphate-buffered saline (PBS) solution, and the increase in Nb content led to better corrosion resistance and reduced water contact angle [7]. Zr65−xTixCu20Al10Fe5 (x = 4, 6) showed better corrosion resistance than Zr65−xTixCu20Al10Fe5 (x = 0, 2) in PBS solution [8]. Zr60+xTi2.5Al10Fe12.5−xCu10Ag5 (x = 0, 2.5, 5) exhibited good corrosion resistance in PBS solution, and higher Zr content resulted in better corrosion resistance due to higher Zr/Al ratio in the passive film [9]. Zr46(Cu4.5/5.5Ag1/5.5)46Al8 showed better biocompatibility and better corrosion resistance than Zr51.9Cu23.3Ni10.5Al14.3 and Zr51Cu25Ni10Al9Ti5 in Hank solution due to the amorphous structure and the formation of Al2O3-enriched passive film [10]. Zr53Al16Co26Pt5 showed the best corrosion resistance, followed by Zr53Al16Co26Pd5, and Zr53Al16Co26Au5 showed the worst corrosion resistance in PBS solution [11]. Zr55.8Al19.4(Co1−xCux)24.8 (x = 0–0.8) were prepared and Zr55.8Al19.4(Co1−xCux)24.8 (x = 0.3) illustrated good corrosion resistance in PBS solution, the largest 12 mm casting diameter, and good mechanical properties [12]. Zr60.5Hf3Al9Fe4.5Cu23 and Zr58.6Al15.4Co18.2Cu7.8 illustrated good corrosion resistance in PBS solution, and good antibacterial properties [13,14]. Zr40Ti37Co12Ni11, Zr50Ti32Cu13Ag5, Zr46Ti40Ag14 and Zr46Ti43Al11 displayed better corrosion resistance than Ti and 316L steel in PBS solution, and Zr46Ti40Ag14 showed better antibacterial properties than Zr46Ti43Al11 [15,16]. Zr45Ti36Fe11Al8 exhibited better corrosion resistance than Ti in PBS solution due to the amorphous structure and the stable ZrO2 and TiO2 film [17]. Compared with Zr65Cu18Al10Ni7, Zr55Cu30Al10Ni5 displayed higher microhardness, higher Ecorr and smaller Epit − Ecorr in PBS solution, and better wear resistance both in air and in PBS solution [18]. Zr62Cu22Al10Fe5Dy1 showed amorphous structure, good corrosion resistance in PBS solution, artificial saliva solution (ASS), Hank solution, and artificial blood plasma (ABP) solution, as well as good biocompatibility [19]. Zr37Co34Cu20Ti9 displayed good corrosion resistance in PBS, ASS, ABP, and Hank solutions, and good biocompatibility [20].
The corrosion behaviours of Zr-based BMGs were influenced by annealing temperature and time. The as-cast Zr56Co28Al16 sample and the samples annealed at 660 K (<Tg), 800 K and 973 K (>Tx) for 5 h displayed poorer corrosion resistance in Ringer’s solution and poorer biocompatibility with higher annealing temperature [21]. Zr58Nb3Cu16Ni13Al10 samples were annealed at 523 K and 673 K (<Tg), as well as at 773 K and 873 K (>Tx), for 6 h, and with higher annealing temperature, the microhardness gradually increased, and the corrosion resistance in 1 mol/L H2SO4 solution at 333 K gradually decreased [22]. Zr60Cu20Ni8Al7Hf3Ti2 samples were annealed at 500 K and 600 K (<Tg) and remained amorphous, and with higher annealing temperature, the microhardness rose, and the corrosion resistance in 0.01 mol/L H2SO4 solution slightly dropped [23]. Zr50.7Ni28Cu9Al12.3 samples were annealed at 719 K (Tg~Tx), 768 K (>Tx), and 810 K for 30 min, and the structures were amorphous, nanocrystals of ZrO2 in an amorphous matrix, and crystals of ZrO2, Al2Zr, Cu10Zr7, and CuZr2 in an amorphous matrix, respectively, and the sample annealed at 768 K exhibited the highest corrosion resistance in 0.5 mol/L H2SO4, 1 mol/L NaCl and 1 mol/L HCl solutions [24]. Zr41.2Cu12.5Ni10Ti13.8Be22.5 and Zr57Cu15.4Ni12.6Al10Nb5 samples annealed at 0.9Tg for 4 h showed better corrosion resistance in NaCl solution than the as-cast sample, due to the decreased free volume [25]. Zr68Al8Ni8Cu16 samples annealed at 673 K and 713 K displayed crystals of Zr2Ni and Zr2Cu in the amorphous matrix with a crystallinity of 10% and 77%, and the sample annealed at 713 K exhibited the maximum microhardness, the lowest strength, the smallest plasticity, and the worst corrosion resistance in 1 mol/L HCl solution [26]. The Zr65Cu15Ni12.5Al7.5 samples were annealed at 643 K, 663 K and 683 K (Tg~Tx) for 10 min, and with higher annealing temperature, the plasticity gradually decreased, and the corrosion resistance in 3.5% NaCl solution gradually decreased, and the cast sample showed higher corrosion resistance than the annealed samples [27]. The fully crystallised Zr48Cu46.5Al4Nb1.5 sample obtained by flash-annealing showed better corrosion resistance than the as-cast Zr48Cu46.5Al4Nb1.5 sample and Zr47.5Cu47.5Al5 sample in 3.5% NaCl solution and 0.05 mol/L H2SO4 solution [28]. Zr56Cu19Ni11Al9Nb5 as-cast sample and samples annealed at 623 K (<Tg), 723 K (Tg~Tx), 823 K and 923 K (>Tx) for 30 min displayed good corrosion resistance in PBS solution, and at higher annealing temperature, the corrosion resistance firstly rose and then dropped, with the best corrosion resistance for the sample annealed at 823 K and the worst corrosion resistance for the sample annealed at 923 K [29].
Zr60Cu20Al10Fe5Ti5 as-cast sample and the samples annealed at 648 K (slightly < Tg) for 1 min and 5 min showed amorphous structure and the sample annealed at 873 K (>Tx) for 1 min displayed crystalline structure with crystals of Zr2Cu and Al2Zr3. They exhibited gradually increased hardness, higher Ecorr and better corrosion resistance in Hank solution, due to structural relaxation, reduced free volume and increased crystallisation [30]. The Zr65Cu17.5Fe10Al7.5 samples were annealed at 573 K (<Tg) for 0.5–4 h, and with prolonged annealing time, the microhardness firstly rose and then dropped, and the sample annealed at 573 K for 1 h showed the maximum microhardness of 487 HV, the highest plasticity of 7.1%, and the highest corrosion resistance in 3.5% NaCl solution due to the decreased free volume [31]. Zr35Ti30Be26.75Cu8.25 samples were annealed at 683 K (Tg~Tx) for 5 min and 12 min, and the corrosion resistance in 3.5% NaCl solution decreased, and the corrosion resistance increased after holding at 77 K for 30 min, and decreased after cryogenic cycling [32]. The as-cast Zr50.7Cu28Ni9Al12.3 sample and the samples annealed at 736 K (Tg~Tx) for 140, 164, and 193 min displayed amorphous structure and crystals in the amorphous matrix with the volume fraction of the crystalline phases of 14%, 40%, and 100%, and the sample annealed for 140 min exhibited the best corrosion resistance in the simulated groundwater, and with longer annealing time, the corrosion resistance gradually dropped [33]. The as-cast Zr59Ti6Cu17.5Fe10Al7.5 sample and the samples annealed at 573 K (<Tg) for 0.5–2 h displayed amorphous structure, and the sample annealed at 573 K for 4 h exhibited crystals of Al3Zr2 in amorphous matrix, and with prolonged annealing time, the strength and plasticity firstly rose and then dropped, with the highest strength and the largest plasticity for the sample annealed at 573 K for 0.5 h, and the corrosion resistance of all the samples remained good in PBS solution [34].
The effects of microstructure, annealing temperature and time displayed complex effects on the corrosion resistance of Zr-based BMGs and BMGCs in simulated bio-environments, and further study is necessary. In this work, the microstructures of as-cast Zr56Cu19Ni11Al9Nb5 metallic glass and the samples annealed at 923 K (>Tx) for 5–30 min were studied by X-ray diffraction (XRD) and scanning electron microscopy (SEM), their microhardness values were tested, and their corrosion behaviours in Hank solution were studied. Our work helps the further understanding of corrosion mechanisms of Zr-based metallic glass and composites and further research and development in biomedical applications.

2. Materials and Methods

2.1. Material Preparation

Zr56Cu19Ni11Al9Nb5 samples were obtained through arc-melting of pure Zr, Cu, Ni, Al and Nb [29]. The samples were cut to 5 mm × 4 mm × 1 mm, as shown in Figure 1, and were annealed at 923 K (far above Tx) for 5, 8, 10, 15, and 30 min. The surfaces were ground using 800–1500 grit sandpaper, polished using 2.5 and 0.5 µm diamond paste, and then cleaned in acetone. The surface roughness of the sample was determined by Dimension IconIR atomic force microscopy (Bruker, MA, USA), as shown in Figure 2, with Ra around 0.007 μm and Rq around 0.014 μm.

2.2. Tests

XRD analysis was performed using a D8 Advance X-ray diffractor (Bruker, Karlsruhe, Germany) with a diffraction angle 2θ in the range of 20–90°, and an SEM image was obtained using Zeiss Ultra Plus field emission scanning electron microscopy (Zeiss, Jena, Germany) with the voltage of 5 kV, in order to investigate the microstructures of the as-cast sample and the samples annealed at 923 K for 5–30 min. The amorphous-to-crystalline ratio is equal to the area of the amorphous structure divided by the area of the crystalline structure. Vickers microhardness measurements were carried out using a MICRO-586 microhardness tester (Shandong Shancai Testing Instrument Co., Ltd., Yantai, Shandong, China) with a load of 2 N and a holding time of 10 s, and the measurements were repeated 10 times for each sample. Potentiodynamic polarisation (PP) tests and electrochemical impedance spectroscopy (EIS) tests of the as-cast and annealed samples in Hank solution were carried out through the CHI 660E electrochemical station (Shanghai CH Instruments, Shanghai, China) [29]. The composition of Hank solution is as follows, 0.14 g/L CaCl2, 0.1 g/L MgCl2·6H2O, 0.1 g/L MgSO4·7H2O, 0.35 g/L NaHCO3, 1 g/L D-glucose, 0.01 g/L phenol red. The saturated calomel electrode (SCE) was used as the reference electrode, and the graphite electrode was used as the counter electrode. The electrodes were stabilised in Hank solution at open circuit potential (OCP) for 60 min. In PP tests, the potential was scanned from −0.8 VSCE (i.e., −0.559 VSHE) to 0.5 VSCE (i.e., 0.741 VSHE) at 0.33 mV/s. In EIS tests, the complex impedance was obtained at OCP, with the frequency of 10−2–105 Hz and the amplitude of 5 mV. Zsimpwin 3.6 software was used to fit the EIS data through the appropriate equivalent circuit (EC). After PP and EIS tests, the corrosion morphology and products were analysed by SH11/YF-III optical microscopy (OM, Shanghai Optical Instrument Factory, Shanghai, China), Zeiss Ultra Plus field emission SEM (Zeiss, Jena, Germany), X-Max 50 X energy dispersive X-ray spectroscopy (EDS, Oxford Instruments, Abingdon, UK) and Thermo Scientific ESCALAB 250Xi X-ray photoelectron spectroscopy (XPS, Thermofisher Scientific, Waltham, MA, USA) [29]. The measurement accuracy of the equipment is shown in Table 1.

3. Results and Discussion

3.1. Microstructure Analysis

The XRD patterns and SEM images of the as-cast Zr56Cu19Ni11Al9Nb5 sample and the samples annealed at 923 K for 5–30 min are shown in Figure 3 and Figure 4. The as-cast sample shows an amorphous structure, and crystallisation occurred in the samples annealed at 923 K for 5–30 min with crystals of Zr2Cu and Zr2Ni. With longer annealing time, the intensity of the diffraction peaks gradually increased, and the crystallisation degree gradually increased, with crystallisation percentages of 0, 17%, 18%, 27%, 27% and 85% for the as-cast sample and the samples annealed at 923 K for 5 min, 8 min, 10 min, 15 min and 30 min, respectively. SEM images display the amorphous structure of the as-cast sample and microstructures of nanocrystals in the amorphous matrix of the annealed samples. With prolonged annealing time, the number of crystals firstly increased and then decreased, and the size of crystals gradually increased, due to the increase in crystallisation degree and the growth and coalescence of crystals. The amorphous-to-crystalline ratio was 4.9, 4.6, 2.7, 2.7 and 0.2 for the samples annealed at 923 K for 5 min, 8 min, 10 min, 15 min and 30 min, respectively.

3.2. Microhardness

Figure 5 shows the microhardness of the as-cast Zr56Cu19Ni11Al9Nb5 sample and the samples annealed at 923 K for 5–30 min. With a longer annealing time, the microhardness gradually rose from 4.89 GPa (i.e., 499 Hv) to 6.37 GPa (i.e., 650 Hv) and then levelled off. The brittleness gradually increased, reaching the maximum, and then slightly dropped. It is due to the increase in crystallisation degree with prolonged annealing time, as shown by XRD analysis and SEM images. The interface between the crystals and the amorphous matrix and the grain boundaries hindered the deformation, leading to the increase in microhardness. A similar trend was reported for Zr65Cu17.5Fe10Al7.5 samples annealed at 573 K (<Tg) for 0.5–4 h, and the microhardness increased from 422 Hv for the as-cast sample to 487 Hv for the sample annealed at 573 K for 1 h and then dropped to 444 Hv for the sample annealed at 573 K for 4 h [31].

3.3. Corrosion Behaviours

Figure 6 shows the PP curves for the as-cast Zr56Cu19Ni11Al9Nb5 sample and the samples annealed at 923 K for 5–30 min in Hank solution, and Table 2 summarises the obtained electrochemical parameters. Figure 7 illustrates the change of corrosion potential Ecorr, corrosion current density icorr, pitting potential Epit, pitting current density ipit with annealing time. For the as-cast sample and the samples annealed at 923 K for 5–10 min, Epit was determined from the kink where the current density increased sharply. For the samples annealed at 923 K for 15 and 30 min, Epit was determined from the intersection point of the tangent lines where the current density increased dramatically. Passivation occurred for all the samples, indicating good corrosion resistance in Hank solution. With longer annealing time, Ecorr and Epit first rose and then dropped, icorr and ipit first dropped and then rose, and the corrosion resistance first rose and then dropped. The sample annealed at 923 K for 5 min exhibited the best corrosion resistance, with high Ecorr at −0.007 VSCE (i.e., 0.234 VSHE), the smallest icorr at 2.20 × 10−7 A·cm−2, the highest Epit at 0.415 VSCE (i.e., 0.656 VSHE), and the largest passivation region Epit − Ecorr at 0.421 VSHE. The sample annealed at 923 K for 30 min exhibited the worst corrosion resistance, with low Ecorr at −0.069 VSCE (i.e., 0.172 VSHE), large icorr at 1.32 × 10−6 A·cm−2, low Epit at −0.001 VSCE (i.e., 0.240 VSHE) and small Epit − Ecorr at 0.068 VSHE. Figure 8 illustrates the Nyquist plots, Bode plots and the equivalent circuit diagram for the EIS results of the as-cast Zr56Cu19Ni11Al9Nb5 sample and the samples annealed at 923 K for 5–30 min in Hank solution, and Table 3 summarises the obtained electrochemical parameters.
The Nyquist plots displayed a half-circle, indicating that the control step was the electrochemical reaction accompanied by the transfer of electrons. The Bode plots showed two time constants, with frequencies around 3 Hz and 1000 Hz. The maximum phase was reached around 3 Hz. The equivalent circuit diagram as shown in Figure 8c was previously used for the fitting of EIS results of Zr45Ti36Fe11Al8 in PBS solution [16], Zr58Nb3Cu16Ni13Al10 as-cast sample and samples annealed at 523 K, 673 K, 773 K and 873 K in 1 mol/L H2SO4 solution at 333 K [21], Zr52Cu32Al10Ni6 in 0.05–0.5 mol/L NaF solutions [35], Al86Ni9Y5 as-spun sample and the samples annealed at 423 K for 5 min and 30 min, at 543 K for 5 min, and at 673 K for 5 min in 0.6 mol/L NaCl [36]. The passivation film contains defects. Rs represents the solution resistance between the working electrode and the reference electrode. Rf represents the film resistance, and Rct represents the charge transfer resistance at the interface between the solution and the film. CPE1 and CPE2 represent the constant phase element (CPE) of the film and the capacitance of the double layer between the solution and the film. The impedance Z = Y 0 1 ω n cos n π 2 j sin n π 2 . Y0 is the constant, ω is the angular frequency, and n is the parameter between 0 and 1. Figure 9 shows the change of the sum of charge transfer resistance and film resistance Rct + Rf with annealing time. With the increase in annealing time, the arc radius gradually increased and then decreased, Rct + Rf gradually decreased and then increased, indicating that the corrosion resistance in Hank solution first increased and then decreased. The sample annealed at 923 K for 5 min showed the largest arc radius and the largest Rct + Rf, 15,489 Ω·cm2, indicating the best corrosion resistance. The sample annealed at 923 K for 30 min displayed the smallest arc radius and the smallest Rct + Rf, 4070 Ω·cm2, indicating the worst corrosion resistance.
Figure 10 and Figure 11 exhibit OM images, SEM images and EDS analysis of the as-cast Zr56Cu19Ni11Al9Nb5 sample and the samples annealed at 923 K for 5–30 min after PP tests in Hank solution. Pitting was observed in all the samples. Spots A, C, E, G, L and N are located in the non-corroded area, with a low oxygen content of 11–48 at.%, and concentrations of Zr, Cu, Ni, Al and Nb of 24–47 at.%, 10–15 at.%, 6–14 at.%, 7–9 at.%, 4–5 at.%, which are similar to the original 56 at.%, 19 at.%, 11 at.%, 9 at.%, and 5 at.%, respectively. Spots B, D, F, M and O are located in the corroded area, with a higher oxygen content of 25–66 at.% and concentrations of Zr, Cu, Ni, Al and Nb of 10–37 at.%, 9–25 at.%, 1–12 at.%, 2–10 at.%, 1–6 at.%, respectively, much lower than the original concentrations. The decrease in Zr content is the largest.
The as-cast Zr56Cu19Ni11Al9Nb5 sample exhibited an amorphous structure, resulting in good corrosion resistance in Hank solution. The annealed sample at 923 K (>Tx) for 5 min displayed a homogeneous microstructure of nanocrystals of Zr2Cu and Zr2Ni in the amorphous matrix, as confirmed by XRD analysis and SEM image. The improved stability of the passivation film led to higher corrosion resistance. With a longer annealing time, the crystallisation content rose. The microstructure became inhomogeneous, with larger crystals of Zr2Cu and Zr2Ni in the amorphous matrix, leading to a drop in corrosion resistance.
Figure 12 shows the XPS analysis of the Zr56Cu19Ni11Al9Nb5 as-cast sample and the samples annealed at 923 K for 5–30 min after EIS tests in Hank solution. The experimental data, the base line, the fitted peaks and the sum of fitted peaks were illustrated in different colours. The peaks at 182.2 eV and 184.6 eV exhibited Zr4+ 3d5/2 and Zr4+ 3d3/2 [3,29,37]. The peaks at 932.0 eV and 952.0 eV represented Cu+ 2p3/2 and Cu+ 2p1/2 [3,29,37]. The peaks at 75.6 eV and 77.6 eV indicated Cu2+ 2p3/2 and Cu2+ 2p1/2 [3,29,37]. The peaks at 851.9 eV and 856.0 eV indicated Ni0 2p3/2 and Ni2+ 2p3/2, and the peaks at 869.5 eV and 873.5 eV indicated Ni0 2p1/2 and Ni2+ 2p1/2 [3,29]. The peak at 74.0 eV represented Al3+ 2p [29,37]. The peaks at 207.0 eV and 209.8 eV indicated Nb5+ 3d3/2 and Nb5+ 3d1/2 [3,29]. The peaks at 529.8 eV and 531.3 eV represented O2− and OH [3,29,37]. Therefore, the corrosion products consist of ZrO2, CuO, Cu2O, Ni(OH)2, Al2O3, and Nb2O5.
The electrode potential of Al/Al3+, Zr/Zr4+, Nb/Nb5+, Ni/Ni2+, Cu/Cu2+, and Cu/Cu+ is as follows, −1.662 VSHE, −1.529 VSHE, −1.200 VSHE, −0.250 VSHE, 0.337 VVSH, and 0.521 VSHE, respectively. Al, Zr, and Nb are more active than Ni and Cu, due to lower potential, and the following corrosion reactions occur first. 2Al + 3H2O → Al2O3 + 3H2, Zr + 2H2O → ZrO2 + 2H2, 2Nb + 5H2O → Nb2O5 + 5H2. The corrosion of Ni and Cu occurs next, with higher potential. Ni + 2H2O → Ni(OH)2 + H2, Cu + H2O → CuO + H2, 2Cu + H2O → Cu2O + H2.
Larsson et al. reported that for the Zr59.3Cu28.8Al10.4Nb1.5 sample manufactured by selective laser melting, the ion release under both simulated physiological and inflammatory conditions is as follows, Zr > Al > Cu > Nb due to the lower potential of Zr and Al, and the ion release under inflammatory condition is higher than that under simulated physiological condition [38]. Zr is a biocompatible element, and Al and Cu are trace elements present in the human body. Ni is a toxic element, which may cause an allergy response. Large ion release higher than the legal limits may cause health problems [39,40,41].

4. Conclusions

The as-cast Zr56Cu19Ni11Al9Nb5 sample showed an amorphous structure, and the samples annealed at 923 K for 5–30 min showed crystals of Zr2Cu and Zr2Ni in the amorphous matrix. With prolonged annealing time, the crystallisation degree gradually increased, and the microhardness gradually increased and then levelled off. Passivation was observed for all the samples in Hank solution, indicating good corrosion resistance. With a longer annealing time, the corrosion resistance first rose and then dropped. The annealed sample at 923 K for 5 min displayed the highest corrosion resistance, with high Ecorr at −0.007 VSCE (i.e., 0.234 VSHE), the smallest icorr at 2.20 × 10−7 A·cm−2, the highest Epit at 0.415 VSCE (i.e., 0.656 VSHE), the largest Epit − Ecorr at 0.421 VSHE, the largest arc radius, and the largest Rct + Rf at 15,489 Ω·cm2. The annealed sample at 923 K for 30 min exhibited the lowest corrosion resistance, with low Ecorr at −0.069 VSCE (i.e., 0.172 VSHE), large icorr at 1.32 × 10−6 A·cm−2 (i.e., 500% increase), low Epit at −0.001 VSCE (i.e., 0.240 VSHE), small Epit − Ecorr at 0.068 VSHE (i.e., 84% decrease), the smallest arc radius, and the smallest Rct + Rf at 4070 Ω·cm2 (i.e., 74% decrease). When the annealing time was appropriate, the microstructure was homogeneous with nanocrystals in an amorphous matrix, resulting in improved passivation film and the rise of corrosion resistance. However, if the annealing time was prolonged, the microstructure became inhomogeneous, with larger crystals in an amorphous matrix, leading to a drop in corrosion resistance. Localised corrosion was observed, with corrosion products of ZrO2, Cu2O, CuO, Ni(OH)2, Al2O3, and Nb2O5. Zr56Cu19Ni11Al9Nb5 sample annealed at 923 K for 5 min exhibited potential candidate for biomedical applications due to high hardness and good corrosion resistance in Hank solution. In the future, in vivo tests can be performed to investigate the biocompatibility.

Author Contributions

Conceptualisation, Z.Z. (Zhiying Zhang); methodology, Z.Z. (Zhiying Zhang); formal analysis, Z.Z. (Zhiying Zhang), J.Z.; investigation, J.Z., X.J., C.Y., Z.Z. (Zikai Zhou) and H.L.; resources, K.W., J.G. and Q.Z.; writing—original draft, Z.Z. (Zhiying Zhang) and J.Z.; writing—review and editing, Z.Z. (Zhiying Zhang), J.Z., K.W., J.G., Q.Z., X.J., C.Y., Z.Z. (Zikai Zhou) and H. Liu; supervision, Z.Z. (Zhiying Zhang); funding acquisition, Z.Z. (Zhiying Zhang). All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (grant number 51502229) and the Start-up Research Foundation of Wuhan University of Technology (grant number 101-40120189).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data shown in this study are available on request from the corresponding author. The data are not publicly available due to privacy reasons.

Acknowledgments

Junwei Wu provided help with the preparations of samples.

Conflicts of Interest

Author Jinguo Gao was employed by the company Hubei Hongle Cable Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Abbreviations

Nomenclature
ABPartificial blood plasma solution
ASSartificial saliva solution
BMGsBulk metallic glasses
BMGCsbulk metallic glass composites
CPEconstant phase element
ECequivalent circuit
EDSenergy dispersive X-ray spectroscopy
EISelectrochemical impedance spectroscopy
OCPopen circuit potential
PBSphosphate-buffered saline
PPPotentiodynamic polarisation
XPSX-ray photoelectron spectroscopy
SCESaturated calomel electrode
SEMscanning electron microscopy
XRDX-ray diffraction
Ecorrcorrosion potential
Epitpitting potential
Epit − Ecorrpassivation region
ipitpitting current density
icorrcorrosion current density
Rctcharge transfer resistance
Rffilm resistance
Rsthe solution resistance between the working electrode and the reference electrode
ωangular frequency

References

  1. Lin, J.G.; Xu, J.; Wang, W.W.; Li, W. Electrochemical behavior of partially crystallized amorphous Al86Ni9La5 alloys. Mater. Sci. Eng. B 2011, 176, 49–52. [Google Scholar] [CrossRef]
  2. Saha, S.K.; Maria, K.H.; Hossain, E.; Bhuiyan, M.A.; Saha, D.K.; Choudhury, S. Impact of annealing time on the formation of nanocrystal in amorphous Fe75.5Si3.5Cu1Nb1B9 alloy and their ultra-soft magnetic properties. J. Bangladesh Acad. Sci. 2016, 40, 137–145. [Google Scholar] [CrossRef]
  3. Yang, Y.J.; Fan, X.D.; Wang, F.L.; Qi, H.N.; Ma, M.Z.; Zhang, X.Y.; Li, G.; Liu, R.P. Effect of Nb content on corrosion behavior of Ti-based bulk metallic glass composites in different solutions. Appl. Surf. Sci. 2019, 471, 108–117. [Google Scholar] [CrossRef]
  4. Imai, K.; Zhou, X.; Liu, X. Application of Zr and Ti-based bulk metallic glasses for orthopaedic and dental device materials. Metals 2020, 10, 203. [Google Scholar] [CrossRef]
  5. Sun, K.; Fu, R.; Liu, X.W.; Xu, L.M.; Wang, G.; Chen, S.Y.; Zhai, Q.J.; Pauly, S. Osteogenesis and angiogenesis of a bulk metallic glass for biomedical implants. Bioact. Mater. 2022, 8, 253–266. [Google Scholar] [CrossRef] [PubMed]
  6. Lin, C.H.; Huang, C.H.; Chuang, J.F.; Lee, H.C.; Liu, M.C.; Du, X.H.; Huang, J.C.; Jang, J.S.C.; Chen, C.H. Simulated body-fluid tests and electrochemical investigations on biocompatibility of metalic glasses. Mater. Sci. Eng. C 2012, 32, 2578–2582. [Google Scholar] [CrossRef]
  7. Lu, X.Y.; Huang, L.; Pang, S.J.; Zhang, T. Formation and biocorrosion behavior of Zr-Al-Co-Nb bulk metallic glasses. Chin. Sci. Bull. 2012, 57, 1723–1727. [Google Scholar] [CrossRef]
  8. Shi, H.; Zhao, W.; Wei, X.; Ding, Y.; Shen, X.; Liu, W. Effect of Ti addition on mechanical properties and corrosion resistance of Ni-free Zr-based bulk metallic glasses for potential biomedical applications. J. Alloys Compd. 2020, 815, 152636. [Google Scholar] [CrossRef]
  9. Hua, N.B.; Huang, L.; Chen, W.Z.; He, W.; Zhang, T. Biocompatibility Ni-free Zr-based bulk metallic glasses with high-Zr-content: Compositional optimization for potential biomedical applications. Mater. Sci. Eng. C 2014, 44, 400–410. [Google Scholar] [CrossRef]
  10. Sun, Y.; Huang, Y.J.; Fan, H.B.; Wang, Y.M.; Ning, Z.L.; Liu, F.Y.; Feng, D.F.; Jin, X.X.; Shen, J.; Sun, J.F.; et al. In vitro and in vivo biocompability of an Ag-bearing Zr-based bulk metallic glass for potential medical use. J. Non-Cryst. Solids 2015, 419, 82–91. [Google Scholar] [CrossRef]
  11. Hua, N.; Liao, Z.; Chen, W.; Huang, Y.; Zhang, T. Effects of noble elements on the glass-forming ability, mechanical property, electrochemical behavior and tribocorrosion resistance of Ni- and Cu-free Zr-Al-Co bulk metallic glass. J. Alloys Compd. 2017, 725, 403–414. [Google Scholar] [CrossRef]
  12. Han, K.; Qiang, J.; Wang, Y.; Zhao, B.; Haussler, P. Zr55.8Al19.4(Co1−xCux)24.8 (x = 0 − 0.8 at.%) bulk metallic glasses for surgical devices applications. J. Iron Steel Res. Int. 2018, 25, 644–649. [Google Scholar] [CrossRef]
  13. Han, K.; Wang, Y.; Qiang, J.; Jiang, H.; Gu, L. Low-cost Zr-based bulk metallic glasses for biomedical devices applications. J. Non-Cryst. Solids 2019, 520, 119442. [Google Scholar] [CrossRef]
  14. Han, K.; Jiang, H.; Wang, Y.; Qiang, J.; Yu, C. Antimicrobial Zr-based bulk metallic glasses for surgical devices applications. J. Non-Cryst. Solids 2021, 564, 120827. [Google Scholar] [CrossRef]
  15. Jabed, A.; Rahman, Z.U.; Khan, M.M.; Haider, W.; Shabib, I. Combinatorial development and in-vitro characterization of the quaternary Zr-Ti-X-Y (X-Y = Cu-Ag/Co-Ni) metallic glass for prospective bio-implants. Adv. Eng. Mater. 2019, 21, 1900726. [Google Scholar] [CrossRef]
  16. Jabed, A.; Khan, M.M.; Camiller, J.; Greenlee-Wacker, M.; Haider, W.; Shabib, I. Property optimization of Zr-Ti-X (X = Ag, Al) metallic glass via combinatorial development aimed at prospective biodemidcal application. Surf. Coat. Technol. 2019, 372, 278–287. [Google Scholar] [CrossRef]
  17. Khan, M.M.; Shabib, I.; Haider, W. A combinatorially developed Zr-Ti-Fe-Al metallic glass with outstanding corrosion resistance for implantable medical devices. Scr. Mater. 2019, 162, 223–229. [Google Scholar] [CrossRef]
  18. Zhao, G.H.; Aune, R.E.; Mao, H.H.; Espallargas, N. Degradation of Zr-based bulk metallic glasses used in load-bearing implants: A tribocorrosion appraisal. J. Mech. Behav. Biomed. Mater. 2016, 60, 56–67. [Google Scholar] [CrossRef]
  19. Jain, A.; Prabhu, Y.; Gunderov, D.; Narayan, R.L.; Saini, P.; Vincent, S.; Sundha, P.; Bagde, A.D.; Bhatt, J. Structural characterization, biocorrosion and in-vitro investigation on Zr62Cu22Al10Fe5Dy1 metallic glass for bio-implant applications. J. Non-Cryst. Solids 2022, 598, 121928. [Google Scholar] [CrossRef]
  20. Prabhu, Y.; Jain, A.; Vincent, S.; Ryu, W.H.; Park, E.S.; Kumar, R.; Bagde, A.D.; Bhatt, J. Compositional design and in vitro investigation on novel Zr-Co-Cu-Ti metallic glass for biomedical applications. Intermetallics 2022, 150, 107692. [Google Scholar] [CrossRef]
  21. Wang, R.R.; Wang, Y.Y.; Yang, J.; Sun, J.M.; Xiong, L.Y. Influence of heat treatment on the mechanical properties, corrosion behavior, and biocompatibility of Zr56Al16Co28 bulk metallic glass. J. Non-Cryst. Solids 2015, 411, 45–52. [Google Scholar] [CrossRef]
  22. Tang, J.; Yu, L.; Qiao, J.; Wang, Y.; Wang, H.; Duan, M.; Chamas, M. Effect of atomic mobility on the electrochemical properties of a Zr58Nb3Cu16Ni13Al10 bulk metallic glass. Electrochim. Acta 2018, 267, 222–233. [Google Scholar] [CrossRef]
  23. Tang, J.; Wang, Y.; Zhu, Q.; Chamas, M.; Wang, H.; Qiao, J.; Zhu, Y.; Normand, B. Passivation behavior of a Zr60Cu20Ni8Al7Hf3Ti2 bulk metallic glass in sulfuric acid solutions. Int. J. Electrochem. Sci. 2018, 13, 6913–6929. [Google Scholar] [CrossRef]
  24. Ge, W.; Li, B.; Axinte, E.; Zhang, Z.; Shang, C.; Wang, Y. Crystallization and corrosion resistance in different aqueous solutions of Zr50.7Ni28Cu9Al12.3 amorphous alloy and its crystallization counterparts. JOM 2017, 69, 776–783. [Google Scholar] [CrossRef]
  25. Aditya, A.; Wu, H.F.; Arora, H.; Mukherjee, S. Amorphous metallic alloys: Pathways for enhanced wear and corrosion resistance. JOM 2017, 69, 2150–2155. [Google Scholar] [CrossRef]
  26. Hua, N.; Liao, Z.; Wang, Q.; Zhang, L.; Ye, Y.; Brechtl, J.; Liaw, P.K. Effects of crystallization on mechanical behavior and corrosion performance of a ductile Zr68Al8Ni8Cu16 bulk metallic glass. J. Non-Cryst. Solids 2020, 529, 119782. [Google Scholar] [CrossRef]
  27. Zhao, X.; Sun, J.; Yu, M.; Zhang, M.; Liu, F.; Zhang, Y. Effects of heat treatment on the thermal, mechanical and corrosion properties of deformed Zr-based bulk metallic glasses. Mater. Chem. Phys. 2020, 256, 123705. [Google Scholar] [CrossRef]
  28. Han, X.; Kaban, I.; Orava, J.; Das, S.M.; Shtefan, V.; Zimmermann, M.V.; Song, K.; Eckert, J.; Nielsch, K.; Herbig, M. Tailoring microstructure and properties of CuZrAl(Nb) metallic-glass-crystal composites and nanocrystalline alloys obtained by flash-annealing. J. Mater. Sci. Technol. 2024, 193, 253–266. [Google Scholar] [CrossRef]
  29. Zhang, Z.; Zhong, X.; Teng, X.; Huang, Y.; Han, H.; Chen, T.; Zhang, Q.; Yang, X.; Gong, Y. Effect of annealing temperature on electrochemical properties of Zr56Cu19Ni11Al9Nb5 in PBS solution. Materials 2023, 16, 3389. [Google Scholar] [CrossRef]
  30. Gonzalez, S.; Pellicer, E.; Surinach, S.; Baro, M.D.; Sort, J. Mechanical and corrosion behaviour of as-cast and annealed Zr60Cu20Al10Fe5Ti5 bulk metallic glass. Intermetallics 2012, 28, 149–155. [Google Scholar] [CrossRef]
  31. Zhou, M.; Hagos, K.; Huang, H.; Yang, M.; Ma, L. Improved mechanical properties and pitting corrosion resistance of Zr65Cu17.5Fe10Al7.5 bulk metallic glass by isothermal annealing. J. Non-Cryst. Solids 2016, 452, 50–56. [Google Scholar] [CrossRef]
  32. Jabed, A.; Kumar, G. Effect of cryogenic cycling and above-Tg annealing on the corrosion behavior of Zr-Cu-Ti-Be metallic glass. J. Non-Cryst. Solids 2023, 608, 122260. [Google Scholar] [CrossRef]
  33. Zhang, L.; Liu, J.; Huang, Y.J.; Ning, Z.; Fan, H.; Sun, J.F. Enhanced corrosion resistance of a ZrCuNiAl bulk metallic glass in simulated groundwater by minor crystallization. Mater. Charact. 2022, 194, 112317. [Google Scholar] [CrossRef]
  34. Shi, H.; Tang, C.; Zhao, X.; Ding, Y.; Ma, L.; Shen, X. Effect of isothermal annealing on mechanical performance and corrosion resistance of Ni-free Zr59Ti6Cu17.5Fe10Al7.5 bulk metallic glass. J. Non-Cryst. Solids 2020, 537, 120013. [Google Scholar] [CrossRef]
  35. Qiu, Z.; Li, Z.; Fu, H.; Zhang, H.; Zhu, Z.; Wang, A.; Li, H.; Zhang, L.; Zhang, H. Corrosion mechanisms of Zr-based bulk metallic glass in NaF and NaCl solutions. J. Mater. Sci. Technol. 2020, 46, 33–43. [Google Scholar] [CrossRef]
  36. Zhang, L.M.; Zhang, S.D.; Ma, A.L.; Hu, H.X.; Zheng, Y.G.; Yang, B.J.; Wang, J.Q. Thermally induced structural evolution on the corrosion behavior of Al-Ni-Y amorphous alloys. Corros. Sci. 2018, 144, 172–183. [Google Scholar] [CrossRef]
  37. Cao, Q.P.; Peng, S.; Zhao, X.N.; Wang, X.D.; Zhang, D.X.; Jiang, J.Z. Effect of Nb substituion for Cu on glass formation and corrosion behavior of Zr-Cu-Ag-Al-Be bulk metallic glass. J. Alloys Compd. 2016, 683, 22–31. [Google Scholar] [CrossRef]
  38. Larsson, L.; Marattukalam, J.J.; Paschalidou, E.M.; Hjorvarsson, B.; Ferraz, N.; Persson, C. Biocompatibility of a Zr-based metallic glass enabled by additive manufacturing. ACS Appl. Bio Mater. 2022, 5, 5741–5753. [Google Scholar] [CrossRef]
  39. Li, T.H.; Wong, P.C.; Chang, S.F.; Tsai, P.H.; Jang, J.S.C.; Huang, J.C. Biocompatibility study on Ni-free Ti-based and Zr-based bulk metallic glasses. Mater. Sci. Eng. C 2017, 75, 1–6. [Google Scholar] [CrossRef]
  40. Zhang, C.; Li, X.M.; Liu, S.Q.; Liu, H.; Yu, L.J.; Liu, L. 3D printing of Zr-based bulk metallic glasses and components for potential biomedical applications. J. Alloys Compd. 2019, 790, 963–973. [Google Scholar] [CrossRef]
  41. Houghton, O.S.; Schmitt, L.Y.; Klotz, U.E.; Costa, M.B.; Greer, A.L. Pd- and Zr-based bulk metallic glasses for jewellery applications: Scracth, wear and tarnish behaviour. Mater. Sci. Eng. A 2025, 924, 147791. [Google Scholar] [CrossRef]
Figure 1. Zr56Cu19Ni11Al9Nb5 samples, (a) as-cast sample, (b) polished sample.
Figure 1. Zr56Cu19Ni11Al9Nb5 samples, (a) as-cast sample, (b) polished sample.
Materials 18 01132 g001
Figure 2. Surface roughness of Zr56Cu19Ni11Al9Nb5 sample.
Figure 2. Surface roughness of Zr56Cu19Ni11Al9Nb5 sample.
Materials 18 01132 g002
Figure 3. XRD patterns of Zr56Cu19Ni11Al9Nb5 as-cast sample and samples annealed at 923 K for 5–30 min.
Figure 3. XRD patterns of Zr56Cu19Ni11Al9Nb5 as-cast sample and samples annealed at 923 K for 5–30 min.
Materials 18 01132 g003
Figure 4. SEM images of Zr56Cu19Ni11Al9Nb5 as-cast sample, and samples annealed at 923 K for 5–30 min. (a) as-cast (b) 5 min (c) 8 min (d) 10 min (e) 15 min (f) 30 min.
Figure 4. SEM images of Zr56Cu19Ni11Al9Nb5 as-cast sample, and samples annealed at 923 K for 5–30 min. (a) as-cast (b) 5 min (c) 8 min (d) 10 min (e) 15 min (f) 30 min.
Materials 18 01132 g004
Figure 5. Microhardness of Zr56Cu19Ni11Al9Nb5 as-cast sample and samples annealed at 923 K for 5–30 min.
Figure 5. Microhardness of Zr56Cu19Ni11Al9Nb5 as-cast sample and samples annealed at 923 K for 5–30 min.
Materials 18 01132 g005
Figure 6. PP curves of Zr56Cu19Ni11Al9Nb5 as-cast sample and samples annealed at 923 K for 5–30 min in Hank solution.
Figure 6. PP curves of Zr56Cu19Ni11Al9Nb5 as-cast sample and samples annealed at 923 K for 5–30 min in Hank solution.
Materials 18 01132 g006
Figure 7. Change of Ecorr, icorr, Epit, ipit with annealing time for as-cast sample and samples annealed at 923 K for 5–30 min. (a) change of Ecorr and icorr with annealing time, (b) change of Epit and ipit with annealing time.
Figure 7. Change of Ecorr, icorr, Epit, ipit with annealing time for as-cast sample and samples annealed at 923 K for 5–30 min. (a) change of Ecorr and icorr with annealing time, (b) change of Epit and ipit with annealing time.
Materials 18 01132 g007
Figure 8. The EIS results for Zr56Cu19Ni11Al9Nb5 as-cast sample and samples annealed at 923 K for 5–30 min in Hank solution, (a) Nyquist plots, (b) Bode plots, (c) the equivalent circuit diagram.
Figure 8. The EIS results for Zr56Cu19Ni11Al9Nb5 as-cast sample and samples annealed at 923 K for 5–30 min in Hank solution, (a) Nyquist plots, (b) Bode plots, (c) the equivalent circuit diagram.
Materials 18 01132 g008
Figure 9. Change of Rct + Rf with annealing time for Zr56Cu19Ni11Al9Nb5 as-cast sample and samples annealed at 923 K for 5–30 min.
Figure 9. Change of Rct + Rf with annealing time for Zr56Cu19Ni11Al9Nb5 as-cast sample and samples annealed at 923 K for 5–30 min.
Materials 18 01132 g009
Figure 10. OM images of the as-cast and annealed Zr56Cu19Ni11Al9Nb5 samples after PP tests. (a) as-cast (b) 5 min (c) 8 min (d) 10 min (e) 15 min (f) 30 min.
Figure 10. OM images of the as-cast and annealed Zr56Cu19Ni11Al9Nb5 samples after PP tests. (a) as-cast (b) 5 min (c) 8 min (d) 10 min (e) 15 min (f) 30 min.
Materials 18 01132 g010
Figure 11. SEM images and EDS analysis of the as-cast and annealed Zr56Cu19Ni11Al9Nb5 samples after PP tests. (a) SEM, as-cast (b) EDS, spot A (c) EDS, spot B (d) SEM, 5 min (e) EDS, spot C (f) EDS, spot D (g) SEM, 8 min (h) EDS, spot E (i) EDS, spot F (j) SEM, 10 min (k) EDS, spot G (l) EDS, spot H (m) SEM, 15 min (n) EDS, spot L (o) EDS, spot M (p) SEM, 30 min (q) EDS, spot N (r) EDS, spot O.
Figure 11. SEM images and EDS analysis of the as-cast and annealed Zr56Cu19Ni11Al9Nb5 samples after PP tests. (a) SEM, as-cast (b) EDS, spot A (c) EDS, spot B (d) SEM, 5 min (e) EDS, spot C (f) EDS, spot D (g) SEM, 8 min (h) EDS, spot E (i) EDS, spot F (j) SEM, 10 min (k) EDS, spot G (l) EDS, spot H (m) SEM, 15 min (n) EDS, spot L (o) EDS, spot M (p) SEM, 30 min (q) EDS, spot N (r) EDS, spot O.
Materials 18 01132 g011aMaterials 18 01132 g011b
Figure 12. XPS analysis of the as-cast and annealed Zr56Cu19Ni11Al9Nb5 samples after EIS tests. (a) Zr 3d (b) Cu 2p (c) Ni 2p (d) Al 2p and Cu 2p (e) Nb 3d (f) O 1s.
Figure 12. XPS analysis of the as-cast and annealed Zr56Cu19Ni11Al9Nb5 samples after EIS tests. (a) Zr 3d (b) Cu 2p (c) Ni 2p (d) Al 2p and Cu 2p (e) Nb 3d (f) O 1s.
Materials 18 01132 g012aMaterials 18 01132 g012b
Table 1. List of equipment and the measurement accuracy.
Table 1. List of equipment and the measurement accuracy.
EquipmentModelParameterAccuracy
X-ray diffractorD8 Advanceangle10−2°
Electrochemical workstationCHI 660Epotential10−4 V
Electrochemical workstationCHI 660Ecurrent10−8 A
Microhardness testerMICRO-586microhardness10−1 MPa
Table 2. PP results of Zr56Cu19Ni11Al9Nb5 as-cast sample and samples annealed at 923 K for 5–30 min in Hank solution.
Table 2. PP results of Zr56Cu19Ni11Al9Nb5 as-cast sample and samples annealed at 923 K for 5–30 min in Hank solution.
Annealing
Time (min)
Ecorr (VSHE)icorr (A·cm−2)Epit (VSHE)ipit (A/cm2)Epit − Ecorr (VSHE)
00.2393.14 × 10−70.4841.67 × 10−60.245
50.2342.20 × 10−70.6562.81 × 10−60.421
80.2404.40 × 10−70.5843.25 × 10−70.344
100.2155.75 × 10−70.3521.64 × 10−60.138
150.1281.05 × 10−60.1455.11 × 10−60.017
300.1721.32 × 10−60.2408.09 × 10−60.068
Table 3. EIS analysis results for as-cast and annealed samples in Hank solution.
Table 3. EIS analysis results for as-cast and annealed samples in Hank solution.
Annealing
Time
(min)
EIS Fitting Results
Rs
(Ω·cm2)
Y01
−1·sn·/cm2)
n1Rf
(Ω·cm2)
Y02
−1·sn·/cm2)
n2Rct
(Ω·cm2)
Rct + Rf
(Ω·cm2)
0111.05 × 10−40.9444442.68 × 10−40.64464490
5111.19 × 10−40.8915,4601.92 × 10−40.712915,489
8111.19 × 10−40.9013,3001.55 × 10−40.724413,344
10121.02 × 10−40.8812,0501.21 × 10−40.753712,087
15161.15 × 10−40.7980532.60 × 10−50.89238076
30179.17 × 10−50.69479.24 × 10−50.8740234070
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, Z.; Zhou, J.; Wang, K.; Gao, J.; Zhang, Q.; Jiang, X.; Yu, C.; Zhou, Z.; Liu, H. Effect of Annealing Time on Corrosion Behaviours of Zr56Cu19Ni11Al9Nb5 in Hank Solution. Materials 2025, 18, 1132. https://doi.org/10.3390/ma18051132

AMA Style

Zhang Z, Zhou J, Wang K, Gao J, Zhang Q, Jiang X, Yu C, Zhou Z, Liu H. Effect of Annealing Time on Corrosion Behaviours of Zr56Cu19Ni11Al9Nb5 in Hank Solution. Materials. 2025; 18(5):1132. https://doi.org/10.3390/ma18051132

Chicago/Turabian Style

Zhang, Zhiying, Jianling Zhou, Kun Wang, Jinguo Gao, Qinyi Zhang, Xinlei Jiang, Chenhao Yu, Zikai Zhou, and Haonan Liu. 2025. "Effect of Annealing Time on Corrosion Behaviours of Zr56Cu19Ni11Al9Nb5 in Hank Solution" Materials 18, no. 5: 1132. https://doi.org/10.3390/ma18051132

APA Style

Zhang, Z., Zhou, J., Wang, K., Gao, J., Zhang, Q., Jiang, X., Yu, C., Zhou, Z., & Liu, H. (2025). Effect of Annealing Time on Corrosion Behaviours of Zr56Cu19Ni11Al9Nb5 in Hank Solution. Materials, 18(5), 1132. https://doi.org/10.3390/ma18051132

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