Alloy | Mg | Zn | Mn | Ca | Y |
Mg–2Zn–0.1Mn–0.3Ca | Bal. | 1.52 | 0.09 | 0.31 | — |
Mg–2Zn–0.1Mn–0.3Ca–0.1Y | Bal. | 1.65 | 0.10 | 0.29 | 0.09 |
Mg–2Zn–0.1Mn–0.3Ca–0.2Y | Bal. | 1.48 | 0.09 | 0.30 | 0.18 |
Mg–2Zn–0.1Mn–0.3Ca–0.3Y | Bal. | 1.46 | 0.10 | 0.33 | 0.26 |
Cite this article as: | Mingfan Qi, Liangyu Wei, Yuzhao Xu, Jin Wang, Aisen Liu, Bing Hao, and Jicheng Wang, Effect of trace yttrium on the microstructure, mechanical property and corrosion behavior of homogenized Mg–2Zn–0.1Mn–0.3Ca–xY biological magnesium alloy, Int. J. Miner. Metall. Mater., 29(2022), No. 9, pp.1746-1754. https://dx.doi.org/10.1007/s12613-021-2327-x |
Magnesium alloys possess the advantages of light weight, good biodegradability and biocompatibility, and similar mechanical property to bone [1–3]. Thus, they have received extensive attention as biodegradable medical implant materials. However, an important research direction is reducing the corrosion rate of magnesium alloys in the body to allow degradation rate to effectively match the tissue regeneration or healing rate [4]. At present, methods to improve corrosion resistance of magnesium alloys mainly include the following: alloying, high purification, coating, surface modification, and chemical treatment. Among them, alloying is the best method for medical applications of magnesium alloys [5–6].
Recently, the research direction of alloying is mainly to add alloying elements with good biocompatibility, such as Zn, Mn, and Ca. Zn can refine the grain of magnesium alloys and improve their mechanical properties. Moreover, adding 1wt% to 2wt% Zn can conduce to form a passive film on the surface of the magnesium alloys and improve their corrosion resistance [7]. Trace Mn can reduce the adverse effect of impurity elements on the corrosion resistance of magnesium alloys [8]. Ca is an effective grain refiner for magnesium alloys, and can improve alloy mechanics and creep resistance [9].
Rare earth (RE) can refine the grain of magnesium alloys and improve their mechanical and corrosion resistance properties. For example, adding Y and Nd can not only refine the grain of the magnesium alloys, but also change the deformation (slip and twinning) mechanism, thereby improving their toughness [10–11]. Adding RE increases fine grain strengthening, dispersion strengthening, solid solution strengthening, and precipitation hardening; it also improves the strength of magnesium alloys. Meanwhile, the formed oxide passivation film with abundant RE elements can improve the corrosion resistance of magnesium alloys [12–14]. At present, magnesium alloys with RE such as WE43, Mg–Li–(Al)–(RE), and LAE442 are used as biomedical materials. Thus, RE is widely used in the field of biological magnesium alloys.
To further improve the corrosion resistance and mechanical property of biological magnesium alloys, Zn, Mn, Ca, and Y are selected as components. Mg–2Zn–0.1Mn–0.3Ca–xY (x = 0, 0.1, 0.2, 0.3) biological magnesium alloy is designed and prepared. The influence of Y content on the microstructure, mechanical properties, and corrosion behavior of magnesium alloys is analyzed by microstructure observation, room temperature tensile experiment, corrosion immersion experiment, and electrochemical test. Consequently, this research provides theoretical basis and reference for the development of new-type Y-containing biological magnesium alloys.
The raw materials for preparing Mg–2Zn–0.1Mn–0.3Ca–xY alloys include pure magnesium, pure zinc, Mg–30Y, Mg–10Mn, and Mg–20Ca master alloys. The raw materials are placed in a ceramic crucible of a ZGJL0.01-50-4K vacuum melting furnace and evacuated, and then argon is introduced into the furnace. The furnace temperature is increased to 750°C. When the raw materials are completely melted, the melt is electromagnetically stirred for 2 min. Subsequently, the melt was poured into a cylinder-shaped graphite crucible with a preheating temperature of 250°C, and an ingot with a diameter of 50 mm and a height of 150 mm was obtained. Table 1 shows the chemical composition of the prepared alloy. The SX-G18133 energy-saving box-type electric furnace is used to homogenize the prepared alloy at 420°C for 18 h.
Alloy | Mg | Zn | Mn | Ca | Y |
Mg–2Zn–0.1Mn–0.3Ca | Bal. | 1.52 | 0.09 | 0.31 | — |
Mg–2Zn–0.1Mn–0.3Ca–0.1Y | Bal. | 1.65 | 0.10 | 0.29 | 0.09 |
Mg–2Zn–0.1Mn–0.3Ca–0.2Y | Bal. | 1.48 | 0.09 | 0.30 | 0.18 |
Mg–2Zn–0.1Mn–0.3Ca–0.3Y | Bal. | 1.46 | 0.10 | 0.33 | 0.26 |
Samples are cut from the four ingots prepared for microstructure observation. After rough grinding and fine grinding, the samples are chemically polished with 20vol% nitric acid methanol, and then picric acid is used for etching. Here, Leica DM 2500 optical microscope (OM) and ZEISS-SUPRA40 scanning electron microscope (SEM) are used to observe the alloy microstructure. SEM is also used to observe tensile fracture morphology and corrosion morphology. Energy-dispersive X-ray spectroscopy (EDS) of the SEM is adopted to test the chemical composition of the micro area. LEXT OLS4000 laser confocal microscope is used to observe and analyze the 3D morphology of the corrosion surface of the alloy after removing the corrosion products. D/Max-RB X-ray diffraction analyzer (XRD) is used for phase analysis of alloys and corrosion products. MDI Jade software is used to analyze the peak value of the test to confirm the types of precipitated phases and corrosion products.
In accordance with the ASTM E8M–04 standard, the sample is obtained from the ingot and processed into tensile samples. The size of sample is shown in Fig. 1. Room temperature tensile test adopts CMT4105 electronic universal testing machine, and tensile rate is 1 mm/min. The tensile result is the average value of the six samples.
The soaking solution is a Kokubo simulated body fluid (SBF) prepared with reference to the Tadashi method [15]. Its pH value is 7.4, and the experiment is carried out in a 37°C water bath environment. Table 2 shows the composition of SBF and human body fluids.
Solution | Na+ | K+ | Ca2+ | Mg2+ | HCO−3 | Cl− | HPO2−4 | SO2−4 |
Plasma | 142.0 | 5.0 | 2.5 | 1.5 | 27.0 | 103.0 | 1.0 | 0.5 |
Kokubo SBF | 142.0 | 5.0 | 2.5 | 1.5 | 4.2 | 147.8 | 1.0 | 0.5 |
A sample is obtained from the ingot and processed into disc shape with a size of ϕ10 mm × 5 mm, and five parallel samples are obtained from the four alloys. Then, the sample is polished with sandpaper and washed with deionized water, ethanol, and acetone in sequence. Subsequently, an electronic balance is used to weigh the dried sample, and the initial weight is recorded as m0.
Vitro immersion test is conducted in accordance with ASTM G31–72 standard. After immersing for 7 d, the samples were removed and washed by an ultrasonic cleaner with deionized water and ethanol in sequence. When morphology observation and phase analysis of the corrosion products are completed, the samples were washed sequentially with boiling solution (200 g/L Cr2O3+ 10g/L AgNO3), acetone, and ethanol for 5 min to remove the corrosion products. After drying, the mass is weighed on an electronic balance and recorded as m1. The average corrosion rate (Pi, mm/a) is calculated by Eqs. (1) and (2).
Pi=87600×ΔmρMg×A×T | (1) |
Δm=m0−m1 | (2) |
where ρMg is the sample density (g·cm−3), A is the total sample area (cm2), and T is the soaking time (h).
The soaking time of the vitro soaking pH value change experiment is 6 d. In addition, SBF is not replaced during the experiment, and the pH value of SBF is measured every 24 h. The ingot is sampled and processed into a size of 10 mm × 10 mm × 5 mm for electrochemical testing. A three-electrode system Versa STAT 3F electrochemical workstation is used to carry out electrochemical experiment. Kokubo SBF was used as the test solution, and the beaker is placed in a constant temperature water bath at 37°C during the test. At the beginning of the test, open circuit potential is measured for 1200 s until it initially stabilizes. Subsequently, the potentiodynamic polarization curve is measured, where the potential sweep is from −250 to −1.3 V vs. SCE, and the sweep rate is 1 mV/s.
The microstructure of Mg–2Zn–0.1Mn–0.3Ca–xY (x = 0, 0.1, 0.2, 0.3) alloy is shown in Fig. 2. Fig. 3 shows grain size and volume fraction of the second phase. Grain size and second-phase content change with increase in Y content. When the Y content increases from 0wt% to 0.3wt%, the average grain size of Mg–2Zn–0.1Mn–0.3Ca–xY alloy decreases from 310 to 144 μm, and the second-phase volume fraction increases from 0.4% to 6.0%. In addition, the increase in Y content coarsens the grain boundary, the second phase at the grain boundary gradually changes from point precipitation to large-scale interconnection precipitation, and the second-phase precipitates in the grains changes from scattered distribution of small particles to uniform distribution of large particles. As a high melting point element, Y is added to the magnesium alloy to limit diffusion power during alloy solidification. Consequently, Y atoms are enriched in front of solid–liquid interface, and the crystalline core increases, thereby refining the crystal grains [16].
Fig. 4 shows the SEM image of Mg–2Zn–0.1Mn–0.3Ca–xY alloy and the EDS analysis result of the precipitated phase. It shows that the second phase precipitates at grain boundaries of each alloy, and the content increases with increase in Y content. In SEM image of the alloy with Y content less than 0.3wt%, the precipitated phases are bright white, with short rod-like morphology and 10 μm length. The EDS analysis shows that the precipitated phase is mainly Mg2Ca. In SEM image of Mg–2Zn–0.1Mn–0.3Ca–0.3Y alloy, the off-white second-phase precipitates at the grain boundary have a herringbone shape, with lamellar morphology and 20 μm length. The EDS analysis shows that the second-phase composition mainly includes Mg, Zn, and Y elements, which are speculated to be in the Mg3Zn6Y phase.
To further determine the phase composition of Mg–2Zn–0.1Mn–0.3Ca–xY alloy, it is subjected to XRD phase analysis. The results are shown in Fig. 5. The figure shows that when Y element content is less than 0.3wt%, that is, when Zn/Y mass ratio is greater than 6.67, the phase composition of alloy mainly includes α-Mg and Mg2Ca. The content of Y element increases to 0.3%. Diffraction peak of Mg3Zn6Y phase appears in alloy when mass ratio of Zn/Y is 6.67. As a type of high melting point rare earth phase, Mg3Zn6Y phase not only restricts grain growth during alloy solidification, but also results in coarsening of alloy grain boundaries.
Fig. 6 shows the tensile test results of Mg–2Zn–0.1Mn–0.3Ca–xY alloy. It shows that when Y element content increases from 0wt% to 0.3wt%, the yield strength (YS) of the alloy continues to increase, and the ultimate tensile strength (UTS) and elongation initially decrease and then increase. The tensile performance of the alloy is the best when Y content is 0.3wt%. Its UTS, YS, and elongation are 118.5 MPa, 68.6 MPa, and 8.3%, respectively, mainly due to the following aspects: (1) solid solution strengthening caused by Y element solid solution; (2) grain refinement strengthening caused by grain refinement; (3) high hardness, high strength, and dispersion of Mg3Zn6Y phase precipitated in the form of eutectic. The alloy matrix is strengthened, and the movement of dislocations is hindered during stretching.
Fig. 7 shows the SEM images of tensile fracture of the Mg–2Zn–0.1Mn–0.3Ca–xY alloy. The tensile fracture of the alloy without Y has many dimples, as well as a small amount of river patterns, tearing edges, and cleavage planes, and the fracture mode has a quasi-cleavage fracture. The tensile fractures of alloys with Y content of 0.1wt% and 0.2wt% have more river patterns and cleavage planes, and their fracture modes are brittle. More fine and deep dimples can be observed in tensile fracture of the alloy with a Y content of 0.3wt%, and the fracture does not contain cleavage fracture characteristics. The fracture mode is mainly plastic fracture, which is also consistent with result of high elongation of the alloy.
The anodic reaction of magnesium alloy corrosion in SBF is that the magnesium matrix loses electrons to generate Mg2+, and the cathodic reaction is that SBF gains electrons to generate H2 and OH−. Therefore, if the corrosion of magnesium alloy is more severe, then more OH− is obtained, and the pH value is higher. Mg–2Zn–0.1Mn–0.3Ca–xY alloy is immersed in SBF with an initial pH value of 7.40, and the pH of solution changes with time, as shown in Fig. 8. The image shows that within 48 h of immersion, the alloy solution without Y shows the slowest increase in pH value as the immersion time is extended, and the pH value increases from 7.40 to 8.85. The pH value of the alloy solution containing Y increases to above 9.0. Within 48–144 h of immersion, the corrosion solution of Mg–2Zn–0.1Mn–0.3Ca–0.3Y alloy exhibits the slowest increase in pH value from 9.12 to 9.45 as the immersion time increases. The pH value of the solution of three other alloys increases to above 9.5. Hence, when immersed for 144 h, Mg–2Zn–0.1Mn–0.3Ca–0.3Y alloy with a lower pH value is the least corroded.
Fig. 9 shows the potentiodynamic polarization curve of Mg–2Zn–0.1Mn–0.3Ca–xY alloy. Tafel extrapolation method is used to fit the polarization curve, and the relevant electrochemical parameters of each alloy are obtained, as shown in Table 3. The corrosion potential of each alloy is relatively different. Y content increases from 0wt% to 0.3wt%, and corrosion potential of the alloy gradually shifts toward positive pole. Anode part of the polarization curve of each alloy has a slow zone, thereby indicating that a corrosion product film is formed on the surface of these alloys when immersed in SBF. Thus, the corrosion process of the alloy is slowed down. A rapidly increasing inflection point is found behind the anodic polarization curve of the alloy; it is the breakdown potential Ebd. If the potential is higher, then the corrosion product film of the alloy is denser and becomes more difficult to break.
Alloy | Ecorr / V vs. SCE | Icorr / (μA·cm−2) | Ebd / V vs. SCE | Pi / (mm·a−1) | −βc / (mV·dec−1) | βa / (mV·dec−1) | Rp / (kΩ·cm2) |
Mg–2Zn–0.1Mn–0.3Ca | −1.74±0.06 | 61.72±4.31 | −1.58±0.12 | 1.41±0.19 | 340.8±21.12 | 354.8±19.24 | 1.22±0.19 |
Mg–2Zn–0.1Mn–0.3Ca–0.1Y | −1.67±0.04 | 79.54±3.63 | −1.57±0.15 | 1.82±0.15 | 311.1±17.13 | 262.9±17.87 | 0.78±0.15 |
Mg–2Zn–0.1Mn–0.3Ca–0.2Y | −1.64±0.05 | 93.49±7.73 | −1.55±0.20 | 2.14±0.23 | 294.2±21.15 | 236.7±20.01 | 0.61±0.21 |
Mg–2Zn–0.1Mn–0.3Ca–0.3Y | −1.57±0.04 | 58.63±4.56 | −1.47±0.09 | 1.34±0.11 | 350.3±19.21 | 320.1±17.42 | 1.24±0.12 |
Notes: Ecorr—corrosion potential, Icorr—corrosion current density, βa—Tafel curve anode slope, βc—Tafel curve cathode slope, Rp—polarization resistance. |
Table 3 shows that the alloy corrosion current density initially increases and then decreases with increase in Y content. The corrosion current density of Mg–2Zn–0.1Mn–0.3Ca–0.3Y alloy is the smallest, and its value is 58.63 μA·cm−2. In line with the corrosion current density, the corrosion rate Pi of the alloy can be calculated by Eq. (3), as follows [17]:
Pi=22.85Icorr | (3) |
Through calculation, when x = 0, 0.1, 0.2, and 0.3, the corrosion rates of Mg–2Zn–0.1Mn–0.3Ca–xY alloy are 1.41, 1.82, 2.14, and 1.34 mm·a−1, respectively. This finding shows that the corrosion resistance of the alloy is the best when Y content is 0.3wt%. This finding is consistent with the conclusion drawn by pH test.
The polarization resistance can also reflect the corrosion resistance of the alloy to a certain extent. The polarization resistance Rp of the alloy can be calculated by Eq. (4), as follows [18]:
Rp=βaβc2.3(βa+βc)Icorr | (4) |
Generally, if the polarization resistance of the alloy is greater, then the alloy becomes more corrosion-resistant. As shown in Table 3, the Rp value of the Mg–2Zn–0.1Mn–0.3Ca–0.3Y alloy is the largest, indicating that the alloy has the best corrosion resistance.
Fig. 10 lists the corrosion rates of Mg–2Zn–0.1Mn–0.3Ca–xY alloys obtained by in vitro immersion weight loss experiments. As Y content increases from 0wt% to 0.3wt%, the corrosion rate of the alloy in SBF initially increases and then decreases. When Y content is 0.3wt%, the alloy exhibits the lowest corrosion rate (2.82 mm·a−1). In addition, the corrosion rate measured by the weight loss experiment is larger than that by the Tafel curve mainly because the soaking time of weightlessness experiment is longer. Thus, the difference in the corrosion performance of each alloy is enlarged.
The surface corrosion product’s morphology of Mg–2Zn–0.1Mn–0.3Ca–xY alloy after immersing in SBF for 7 d is shown in Fig. 11. The figure shows that flatness of the corrosion surface of each alloy is different, and the surface corrosion product film appears to be cracked in different degrees. The surface film layer of the alloy without Y is relatively flat, but the film layer has a large crack width and is not dense. The surface film layer of the alloy with Y contents of 0.1wt% and 0.2wt% is uneven, and the corrosion products generated are uneven, with the film layer falling off. When Y content is 0.3wt%, the corrosion product film layer of the alloy is relatively flat. At the same time, the corrosion product is dense, and the crack width is small, with a good protective effect on the magnesium matrix. This finding also explains why the Mg–2Zn–0.1Mn–0.3Ca–0.3Y alloy has the best corrosion resistance.
Fig. 12 shows the 3D morphology of the Mg–2Zn–0.1Mn–0.3Ca–xY alloy after being soaked for 7 d and removing the surface corrosion products. It can be seen that the corrosion surface of the alloy without Y is flat, but large and deep pittings are found, indicating that the alloy has localized corrosion. After adding 0.1wt% and 0.2wt% of Y element, the corrosion surface of the alloy becomes highly undulate, and the size of pittings decreases but their number is large. This result indicates that the local corrosion phenomenon is still serious. The corrosion surface of Mg–2Zn–0.1Mn–0.3Ca–0.3Y alloy is generally flat, and its pittings have small size, shallow depth, and small quantity. This finding is mainly attributed to the formation of a fine corrosion product layer on the surface of the alloy after adding 0.3wt% Y, which insulates the contact between the solution and the substrate and inhibits the occurrence of local pitting.
Fig. 13 shows an EDS analysis result of corrosion products on the surface of Mg–2Zn–0.1Mn–0.2Ca–0.2Y alloy after immersion in SBF for 7 d. The main components of corrosion products are Mg, O, Ca, P, and little Y and Na elements. The alloy is immersed in SBF, the corrosion product Mg(OH)2 is superimposed on the surface to form a film, and the Mg element is mainly obtained from Mg(OH)2. The atomic ratio of Ca element to P element in the corrosion product is approximately 1:1. A CaHPO4 compound is formed by the reaction of the Mg2Ca phase in the alloy with the solution. This result indicates that the Mg2Ca phase is heavily corroded in SBF.
The XRD analysis results of the surface of Mg–2Zn–0.1Mn–0.3Ca–xY alloy after corrosion are shown in Fig. 14. The figure shows that for the alloy with Y content less than 0.3wt%, the main residues on the surface after corrosion are Mg(OH)2 and CaHPO4·(2H2O). However, for the alloy with 0.3wt% Y content, only Mg(OH)2 was detected in the surface corrosion products, and no CaHPO4·(2H2O) phase was found. In addition, the Mg3Zn6Y phase is found in the XRD pattern of Mg–2Zn–0.1Mn–0.3Ca–0.3Y alloy. It may be the second phase remaining after the magnesium matrix is corroded, indicating that the magnesium matrix is corroded in the SBF prior to Mg3Zn6Y phase.
The influence of trace Y on the corrosion resistance of Mg–2Zn–0.1Mn–0.3Ca–xY alloy mainly includes potential of the precipitated second phase and film-forming ability of the product formed after corrosion. The corrosion process principle is shown in Fig. 15. The potential of each phase in the Mg–2Zn–0.1Mn–0.3Ca–xY alloy is Mg2Ca < Mg < Mg3Zn6Y [19]. The alloy with Y content less than 0.3wt% contains a large amount of Mg2Ca phase and does not contain Mg3Zn6Y phase. Thus, the Mg2Ca phase is acted as the anode to form microgalvanic corrosion with the substrate. The Mg2Ca phase reacts with SBF to gradually form pittings, resulting in uneven corrosion surface, and the corrosion products contain more CaHPO4·2H2O. The Mg2Ca phase in the Mg–2Zn–0.1Mn–0.3Ca alloy occupies the largest proportion in the second phase. Thus, the pittings on the surface of the corroded alloy are large in size and deep. The Mg–2Zn–0.1Mn–0.3Ca–0.3Y alloy contains Mg3Zn6Y phase as a small cathode. It can form microgalvanic corrosion with the substrate as a large anode, leading to a large-scale uniform corrosion of the magnesium substrate. In addition, the Mg3Zn6Y phase remains on surface of the alloy. This finding is verified in the XRD pattern shown in Fig. 14. In addition, Mg–2Zn–0.1Mn–0.3Ca–0.3Y alloy has the smallest grain size, the densest grain boundary distribution, and the strongest film-forming ability to fix corrosion products, thereby isolating the solution to protect the matrix and obtaining the flattest corrosion morphology.
(1) The addition of rare earth Y can significantly refine the Mg–2Zn–0.1Mn–0.3Ca–xY alloy grains, which promote the formation of second phase. When Y content increases from 0% to 0.3%, the grain size decreases from 310 to 144 μm, and the volume fraction of the second phase increases from 0.4% to 6.0%. When the Y content reaches 0.3wt%, the Mg3Zn6Y phase begins to precipitate in the alloy.
(2) As Y content increases from 0wt% to 0.3wt%, the yield strength of Mg–2Zn–0.1Mn–0.3Ca–xY alloy continues to increase, and the tensile strength and elongation initially decrease and then increase; the tensile properties of the alloy are the best when Y content is 0.3wt%.
(3) When Y content is 0.3wt%, the Mg–2Zn–0.3Ca–0.1Mn–xY alloy exhibits the best corrosion resistance in SBF mainly attributed to the grain refinement and the precipitated Mg3Zn6Y phase.
This work was financially supported by the National Natural Science Foundation of China (Nos. 52005034 and 52027805), the China Postdoctoral Science Foundation Funded Project (No. 2021M691860), the Beijing Postdoctoral Research Foundation (No. 2021-ZZ-073), the Zhuhai Industry-University-Research Cooperation Project (No. ZH22017001200176PWC), and the Tai’an City Science and Technology Innovation Major Project (No. 2021ZDZX011).
The authors declare no potential conflict of interest.
[1] |
J.L. Su, J. Teng, Z.L. Xu, and Y. Li, Biodegradable magnesium-matrix composites: A review, Int. J. Miner. Metall. Mater., 27(2020), No. 6, p. 724. DOI: 10.1007/s12613-020-1987-2
|
[2] |
M. Razzaghi, M. Kasiri-Asgarani, H.R. Bakhsheshi-Rad, and H. Ghayour, In vitro bioactivity and corrosion of PLGA/hardystonite composite-coated magnesium-based nanocomposite for implant applications, Int. J. Miner. Metall. Mater., 28(2021), No. 1, p. 168. DOI: 10.1007/s12613-020-2072-6
|
[3] |
Z. Zhang, J.H. Zhang, J. Wang, Z.H. Li, J.S. Xie, S.J. Liu, K. Guan, and R.Z. Wu, Toward the development of Mg alloys with simultaneously improved strength and ductility by refining grain size via the deformation process, Int. J. Miner. Metall. Mater., 28(2021), No. 1, p. 30. DOI: 10.1007/s12613-020-2190-1
|
[4] |
M.P. Staiger, A.M. Pietak, J. Huadmai, and G. Dias, Magnesium and its alloys as orthopedic biomaterials: A review, Biomaterials, 27(2006), No. 9, p. 1728. DOI: 10.1016/j.biomaterials.2005.10.003
|
[5] |
M.W. Yu, J.Y. Li, J.X. Li, J. Wang, H.Y. Lai, and Y. Zhang, Effects of trace Sr on microstructure, mechanical properties and corrosion resistance of Mg–0.2Zn–0.1Mn–xSr biomaterials, Rare Met. Mater. Eng., 48(2019), No. 12, p. 4016.
|
[6] |
Y.Z. Xu, J.Y. Li, M.F. Qi, L.H. Liao, and Z.J. Gao, Enhanced mechanical properties of Mg–Zn–Y–Zr alloy by low-speed indirect extrusion, J. Mater. Res. Technol., 9(2020), No. 5, p. 9856. DOI: 10.1016/j.jmrt.2020.06.029
|
[7] |
R.Q. Zhang, J.F. Wang, S. Huang, S.J. Liu, and F.S. Pan, Substitution of Ni for Zn on microstructure and mechanical properties of Mg–Gd–Y–Zn–Mn alloy, J. Magnes. Alloys, 5(2017), No. 3, p. 355. DOI: 10.1016/j.jma.2017.07.002
|
[8] |
W.W. He, E.L. Zhang, and K. Yang, Effect of Y on the bio-corrosion behavior of extruded Mg–Zn–Mn alloy in Hank’s solution, Mater. Sci. Eng. C, 30(2010), No. 1, p. 167. DOI: 10.1016/j.msec.2009.09.014
|
[9] |
Z.R. Xie, C. Zhang, H.C. Pan, Y.X. Wang, Y.P. Ren, and G.W. Qin, Microstructures and bio-corrosion resistances of as-extruded Mg–Ca alloys with ultra-fine grain size, Rare Met., 2017, DOI: 10.1007/s12598-017-0945-2.
|
[10] |
L. Zhang, Z. Liu, and P.L. Mao, Effect of annealing on the microstructure and mechanical properties of Mg−2.5Zn−0.5Y alloy, Int. J. Miner. Metall. Mater., 21(2014), No. 8, p. 779. DOI: 10.1007/s12613-014-0971-0
|
[11] |
E. Aghion, G. Levy, and S. Ovadia, In vivo behavior of biodegradable Mg–Nd–Y–Zr–Ca alloy, J. Mater. Sci. -Mater. Med., 23(2012), No. 3, p. 805.
|
[12] |
Y.Z. Xu, J.Y. Li, M.F. Qi, J.B. Gu, and Y. Zhang, Effect of extrusion on the microstructure and corrosion behaviors of biodegradable Mg–Zn–Y–Gd–Zr alloy, J. Mater. Sci., 55(2020), No. 3, p. 1231. DOI: 10.1007/s10853-019-03978-8
|
[13] |
S.Q. Yin, W.C. Duan, W.H. Liu, L. Wu, J.M. Yu, Z.L. Zhao, M. Liu, P. Wang, J.Z. Cui, and Z.Q. Zhang, Influence of specific second phases on corrosion behaviors of Mg−Zn−Gd−Zr alloys, Corros. Sci., 166(2020), art. No. 108419. DOI: 10.1016/j.corsci.2019.108419
|
[14] |
C. Zhang, L. Wu, H. Liu, G.S. Huang, B. Jiang, A. Atrens, and F.S. Pan, Microstructure and corrosion behavior of Mg–Sc binary alloys in 3.5wt% NaCl solution, Corros. Sci., 174(2020), art. No. 108831. DOI: 10.1016/j.corsci.2020.108831
|
[15] |
T. Kokubo and H. Takadama, How useful is SBF in predicting in vivo bone bioactivity? Biomaterials, 27(2006), No. 15, p. 2907. DOI: 10.1016/j.biomaterials.2006.01.017
|
[16] |
D.F. Zhang, X.X. Xu, F.G. Qi, X.X. Guo, and Z.T. Zhu, Research status of yttrium-containing Mg-Zn based magnesium alloys, Foundary, 61(2012), No. 3, p. 266.
|
[17] |
J.X. Li, Y. Zhang, J.Y. Li, and J.X. Xie, Effect of trace HA on microstructure, mechanical properties and corrosion behavior of Mg−2Zn−0.5Sr alloy, J. Mater. Sci. Technol., 34(2018), No. 2, p. 299. DOI: 10.1016/j.jmst.2017.06.013
|
[18] |
Y. Zhang, J.X. Li, and J.Y. Li, Microstructure, mechanical properties, corrosion behavior and film formation mechanism of Mg−Zn−Mn−xNd in Kokubo’s solution, J. Alloys Compd., 730(2018), p. 458. DOI: 10.1016/j.jallcom.2017.09.325
|
[19] |
H.S. Jiang, Microstructure and Mechanical Properties of High Strength Mg−Zn−(Y/Gd)−Zr−(Ca) Alloys Containing W Phase [Dissertation], Harbin Institute of Technology, Harbin, 2017.
|
[1] | Baosheng Liu, Jiali Wei, Shaohua Zhang, Yuezhong Zhang, Pengpeng Wu, Daqing Fang, Guorui Ma. Microstructures, corrosion behavior and mechanical properties of as-cast Mg–6Zn–2X(Fe/Cu/Ni) alloys for plugging tool applications [J]. International Journal of Minerals, Metallurgy and Materials, 2024, 31(4): 697-711. DOI: 10.1007/s12613-023-2775-6 |
[2] | Xiangyang Peng, Yuhai Tang, Xiangbin Ding, Zhichao Lu, Shuo Hou, Jianming Zhou, Shuyin Han, Zhaoping Lü, Guangyao Lu, Yuan Wu. Fe-based amorphous coating prepared using high-velocity oxygen fuel and its corrosion behavior in static lead–bismuth eutectic alloy [J]. International Journal of Minerals, Metallurgy and Materials, 2022, 29(11): 2032-2040. DOI: 10.1007/s12613-022-2420-9 |
[3] | Hassan Jafari, Amir Houshang Mojiri Tehrani, Mahsa Heydari. Effect of extrusion process on microstructure and mechanical and corrosion properties of biodegradable Mg–5Zn–1.5Y magnesium alloy [J]. International Journal of Minerals, Metallurgy and Materials, 2022, 29(3): 490-502. DOI: 10.1007/s12613-021-2275-5 |
[4] | Masoud Sabzi, Sadegh Moeini Far, Saeid Mersagh Dezfuli. Effect of melting temperature on microstructural evolutions, behavior and corrosion morphology of Hadfield austenitic manganese steel in the casting process [J]. International Journal of Minerals, Metallurgy and Materials, 2018, 25(12): 1431-1438. DOI: 10.1007/s12613-018-1697-1 |
[5] | Saeid Mersagh Dezfuli, Ali Shanaghi, Saeid Baghshahi. Effect of Al2O3 and Y2O3 on the corrosion behavior of ZrO2-benzotriazole nanostructured coatings applied on AA2024 via a sol-gel method [J]. International Journal of Minerals, Metallurgy and Materials, 2018, 25(11): 1344-1353. DOI: 10.1007/s12613-018-1688-2 |
[6] | Bao-biao Yu, Hong Yan, Qing-jie Wu, Zhi Hu, Fan-hui Chen. Microstructure and corrosion behavior of Al3Ti/ADC12 composite modified with Sr [J]. International Journal of Minerals, Metallurgy and Materials, 2018, 25(7): 840-848. DOI: 10.1007/s12613-018-1633-4 |
[7] | Mostafa Amirjan, Mansour Bozorg. Properties and corrosion behavior of Al based nanocomposite foams produced by the sintering-dissolution process [J]. International Journal of Minerals, Metallurgy and Materials, 2018, 25(1): 94-101. DOI: 10.1007/s12613-018-1551-5 |
[8] | P. Laxman Mani Kanta, V. C. Srivastava, K. Venkateswarlu, Sharma Paswan, B. Mahato, Goutam Das, K. Sivaprasad, K. Gopala Krishna. Corrosion behavior of ultrafine-grained AA2024 aluminum alloy produced by cryorolling [J]. International Journal of Minerals, Metallurgy and Materials, 2017, 24(11): 1293-1305. DOI: 10.1007/s12613-017-1522-2 |
[9] | Se-fei Yang, Ying Wen, Pan Yi, Kui Xiao, Chao-fang Dong. Effects of chitosan inhibitor on the electrochemical corrosion behavior of 2205 duplex stainless steel [J]. International Journal of Minerals, Metallurgy and Materials, 2017, 24(11): 1260-1266. DOI: 10.1007/s12613-017-1518-y |
[10] | He-rong Zhou, Xiao-gang Li, Chao-fang Dong, Kui Xiao, Tai Li. Corrosion behavior of aluminum alloys in Na2SO4 solution using the scanning electrochemical microscopy technique [J]. International Journal of Minerals, Metallurgy and Materials, 2009, 16(1): 84-88. DOI: 10.1016/S1674-4799(09)60014-5 |
1. | Rahul Mukherjee, Shubhabrata Datta. Materials Informatics Driven Designing Mg Alloys for Biodegradable Short-Term Implants Using Machine Learning. JOM, 2025, 77(1): 20. DOI:10.1007/s11837-024-06924-5 |
2. | Xian-Wen Chen, Hong-Yu Yang, Bai-Xin Dong, et al. The Development of High-strength Flame-retardant Magnesium Alloys. Journal of Materials Research and Technology, 2025. DOI:10.1016/j.jmrt.2025.04.076 |
3. | Wenxi Zhang, Zhangzelong Zhuo, Dan Xu, et al. Superhydrophobic and corrosion-resistant siloxane-modified MgAl–LDHs coatings on magnesium alloy prepared under mild conditions. International Journal of Minerals, Metallurgy and Materials, 2025, 32(2): 450. DOI:10.1007/s12613-024-2927-3 |
4. | Hongxuan Liu, Ting–an Zhang. Development of aqueous magnesium–air batteries: From structure to materials. Journal of Alloys and Compounds, 2024, 988: 174262. DOI:10.1016/j.jallcom.2024.174262 |
5. | Hongxia Li, Wenjun Xu, Yufei Zhang, et al. Prediction of the thermal conductivity of Mg-Al-La alloys by CALPHAD method. International Journal of Minerals, Metallurgy and Materials, 2024, 31(1): 129. DOI:10.1007/s12613-023-2759-6 |
6. | Wen Peng, Zehang Lu, Enyang Liu, et al. Preparation, Mechanical Properties, and Degradation Behavior of Zn-1Fe-xSr Alloys for Biomedical Applications. Journal of Functional Biomaterials, 2024, 15(10): 289. DOI:10.3390/jfb15100289 |
7. | Zheng Tian, Bai-Xin Dong, Xian-Wen Chen, et al. Effects and mechanisms of rare earth and calcium on the flame retardancy of magnesium alloys. Journal of Materials Research and Technology, 2024, 30: 9542. DOI:10.1016/j.jmrt.2024.06.009 |
8. | Hong Yang, Wenlong Xie, Jiangfeng Song, et al. Current progress of research on heat-resistant Mg alloys: A review. International Journal of Minerals, Metallurgy and Materials, 2024, 31(6): 1406. DOI:10.1007/s12613-023-2802-7 |
9. | Mariyappan Mahalingam, Poovazhagan Lakshmanan, Gnanavelbabu Annamalai, et al. Tensile strength and immersion corrosion behavior of nano-ZnO, rare earth Gd reinforced Mg nanocomposites developed by novel stir ultrasonication process. Advanced Composite Materials, 2024, 33(6): 1325. DOI:10.1080/09243046.2024.2347665 |
10. | Bo Wang, Quanzhuo Jin, Xuchao Du, et al. Corrosion mechanism and microstructure evolution of yttrium-doped marine steel. Journal of Materials Research and Technology, 2024, 28: 2752. DOI:10.1016/j.jmrt.2023.12.189 |
11. | Sijia Nie, Zhibin Zheng, Yanxin Qiao, et al. Corrosion behavior of as-cast Al0.75CoCr1.25FeNi high entropy alloy in 0.5 mol/L sulfuric acid. Advanced Composites and Hybrid Materials, 2024, 7(5) DOI:10.1007/s42114-024-00951-1 |
12. | Ce Zhang, Ruifeng Ju, Yunlong Li, et al. Wire-arc additive manufacturing of Mg-Gd-Y-Zn-Zr alloy: Microstructure and mechanical properties. Journal of Materials Research and Technology, 2024, 32: 3083. DOI:10.1016/j.jmrt.2024.08.150 |
13. | Yoshihito Kawamura, Fuka Shimada, Kazuma Hamada, et al. Development of Biomedical Mg–1.0Ca–0.5Zn–0.1Y–0.03Mn (at%) Alloy by Rapidly Solidified Powder Metallurgy Processing. MATERIALS TRANSACTIONS, 2023, 64(9): 2333. DOI:10.2320/matertrans.MT-M2023066 |
14. | Yan Yang, Xiaoming Xiong, Jing Chen, et al. Research advances of magnesium and magnesium alloys worldwide in 2022. Journal of Magnesium and Alloys, 2023, 11(8): 2611. DOI:10.1016/j.jma.2023.07.011 |
15. | Kaiyang Wang, Shaojie Lv, Honghui Wu, et al. Recent research progress on the phase-field model of microstructural evolution during metal solidification. International Journal of Minerals, Metallurgy and Materials, 2023, 30(11): 2095. DOI:10.1007/s12613-023-2710-x |
16. | Jinshu Xie, Zhi Zhang, Shujuan Liu, et al. Designing new low alloyed Mg—RE alloys with high strength and ductility via high-speed extrusion. International Journal of Minerals, Metallurgy and Materials, 2023, 30(1): 82. DOI:10.1007/s12613-022-2472-x |
17. | Runjia Liu, Yan Liu, Qiwen Yong, et al. Highly corrosion-resistant ZIF-8-integrated micro-arc oxidation coating on Mg alloy. Surface and Coatings Technology, 2023, 463: 129505. DOI:10.1016/j.surfcoat.2023.129505 |
Alloy | Mg | Zn | Mn | Ca | Y |
Mg–2Zn–0.1Mn–0.3Ca | Bal. | 1.52 | 0.09 | 0.31 | — |
Mg–2Zn–0.1Mn–0.3Ca–0.1Y | Bal. | 1.65 | 0.10 | 0.29 | 0.09 |
Mg–2Zn–0.1Mn–0.3Ca–0.2Y | Bal. | 1.48 | 0.09 | 0.30 | 0.18 |
Mg–2Zn–0.1Mn–0.3Ca–0.3Y | Bal. | 1.46 | 0.10 | 0.33 | 0.26 |
Solution | Na+ | K+ | Ca2+ | Mg2+ | HCO−3 | Cl− | HPO2−4 | SO2−4 |
Plasma | 142.0 | 5.0 | 2.5 | 1.5 | 27.0 | 103.0 | 1.0 | 0.5 |
Kokubo SBF | 142.0 | 5.0 | 2.5 | 1.5 | 4.2 | 147.8 | 1.0 | 0.5 |
Alloy | Ecorr / V vs. SCE | Icorr / (μA·cm−2) | Ebd / V vs. SCE | Pi / (mm·a−1) | −βc / (mV·dec−1) | βa / (mV·dec−1) | Rp / (kΩ·cm2) |
Mg–2Zn–0.1Mn–0.3Ca | −1.74±0.06 | 61.72±4.31 | −1.58±0.12 | 1.41±0.19 | 340.8±21.12 | 354.8±19.24 | 1.22±0.19 |
Mg–2Zn–0.1Mn–0.3Ca–0.1Y | −1.67±0.04 | 79.54±3.63 | −1.57±0.15 | 1.82±0.15 | 311.1±17.13 | 262.9±17.87 | 0.78±0.15 |
Mg–2Zn–0.1Mn–0.3Ca–0.2Y | −1.64±0.05 | 93.49±7.73 | −1.55±0.20 | 2.14±0.23 | 294.2±21.15 | 236.7±20.01 | 0.61±0.21 |
Mg–2Zn–0.1Mn–0.3Ca–0.3Y | −1.57±0.04 | 58.63±4.56 | −1.47±0.09 | 1.34±0.11 | 350.3±19.21 | 320.1±17.42 | 1.24±0.12 |
Notes: Ecorr—corrosion potential, Icorr—corrosion current density, βa—Tafel curve anode slope, βc—Tafel curve cathode slope, Rp—polarization resistance. |