Microstructure and superior mechanical properties of cast Mg–12Gd–2Y–0.5Sm–0.5Sb–0.5Zr alloy

Ke-jie Li,Quan-an Li
DOI: https://doi.org/10.1016/j.msea.2011.03.049
2011-01-01
Abstract:Highlights ► Mg–Gd–Y alloy exhibits higher mechanical properties at room and high temperatures. ► The experimental alloy exhibits excellent mechanical properties in 20–300 °C. ► The alloy exhibits anomalous temperature dependence of tensile strength in 20–300 °C. ► The high tensile strength was mainly associated with LPO structure β′ precipitates. Abstract The Mg–12Gd–2Y–0.5Sm–0.5Sb–0.5Zr (wt.%) alloy was prepared by casting technology. The microstructure of alloy was then investigated after solution and aging treatment (i.e., T6 heat treatment). Tensile tests were performed at a crosshead speed of 1 mm/min at ambient and elevated temperature atmosphere. The results show that the aged alloy was mainly composed of α-Mg matrix and dispersed long-period ordered β′ precipitates. At 250 °C, the alloy has shown the superior tensile strength (i.e., 345.5 MPa). The remarkable high strength of experimental alloy was mainly associated with solution strengthen of RE and precipitation strengthening of dispersive LPO structure β′ precipitates in Mg matrix. The LPO β′ precipitates provide important strengthening sources in experimental alloy, especially at elevated temperatures. Keywords Mg–12Gd–2Y–0.5Sm–0.5Sb–0.5Zr alloy Microstructure Long-period ordered (LPO) phase Mechanical properties 1 Introduction Mg alloys are attractively as structural light-weight materials in automotive and aerospace industries for their low density, higher specific strength, good cast ability and recoverable. However, compared with Al alloys, the application of the Mg alloys is very limited due to their lower mechanical properties especially at elevated temperatures [1–3] . Thus, it is very significant to develop novel high performance Mg alloys. Mg–RE alloys containing gadolinium element are known to offer excellent heat-resistant mechanical properties by solid solution strengthening and precipitation strengthening. The binary phase diagram of Mg–Gd shows an eutectic with a solubility of 23.5 wt.% Gd at the temperature of 548 °C and 3.8 wt.% Gd at 200 °C, and Mg–Gd based alloy is prone to form supersaturate solid solution after solidification. At the same time, the Mg–Gd based alloy shows remarkable aging hardening due to the precipitation of β′ phase even above 200 °C. The Mg–Gd–Y based alloy exhibits good combination of strength and creep-resistance in the peak aged hardness even better than commercial aluminum alloys, and has the potential application in aerospace industry [4–7] . In this work, the microstructure and mechanical properties of high strength Mg–12Gd–2Y–0.5Sm–0.5Sb–0.5Zr alloy were investigated at ambient and elevated temperature. The morphology and structure of β′ phase has been observed by TEM and HRTEM. These results would have implication for further developing novel applied Mg–Gd–Y based alloy. 2 Experimental procedure An alloy ingot with a nominal composition of Mg–12Gd–2Y–0.5Sm–0.5Sb–0.5Zr (wt.%) was then prepared from Mg–24% Gd, Mg–24% Y, Mg–24% Sm, Mg–30% Zr master alloys, pure Mg and Sb in the induction melting furnace under protective mixed atmosphere of SF 6 (1 vol.%) and CO 2 (bal.). Solution-treatment was carried out at 525 °C for 6 h, and then quenched into water at 80°C. Aging treatment was performed at 225°C for 10 h. The samples were machined to a gauge length of 55 mm and a gauge diameter of 5 mm as shown in Fig. 1 . Tensile tests were performed at a crosshead speed of 1 mm/min at ambient and elevated temperature using AG-I250 kN precision universal material test machine. Microstructure was observed by optical microscope (OM), high resolution transmission electron microscope (HRTEM; JEM-2100) operating at an acceleration voltage of 200 kV. The specimens were mechanically polished and then etched in a solution of 30 ml nitric acid and 70 ml ethanol. Specimens for TEM observation were prepared by a twin-jet technique. Fracture surface and composition of the alloy were investigated using a scanning electron microscopy (SEM; JSM-5610LV) device with energy dispersive spectroscopy (EDS). 3 Results 3.1 Microstructure The optical image of aged alloy is shown in Fig. 2 . The typical aged alloy consisted of α-Mg matrix and partially divorced particles. The average grain size was about 69 μm. Fig. 3 is SEM image of aged alloy sample and corresponding EDS analysis results, which shows that the aged alloy contains different phases, namely: the major phase which is saturated Gd in Mg matrix, Mg 29 Y 33 Gd 10 SmSb (point B) compound and granular Mg 99 Y 12 Gd 20 Sm 3 Sb (point C) compound (both were named as Mg–RE). According to Mg–Gd system phase diagram, the solid solubility limitation of Gd in Mg is less than 3.8 wt.% at ambient temperature, which opposite to the EDS analysis in area A. It should be the result that the nano-size β′ precipitates in the alloy cannot be distinguished by EDS detector. Fig. 4 shows bright field image and corresponding selected area electron diffraction (SAED) pattern from sample with the incident electron beam parallel to [0 0 0 1] α . The oval plate-shape precipitates were observed with three variants. These grew to ∼35 nm in diameter, with a thickness of <15 nm ( Fig. 4 (a)). The SAED pattern ( Fig. 4 (b)) shows extra diffraction spots at 1 / 4 { 1 ¯     0     1     0 } α , 1 / 2 { 1 ¯     0     1     0 } α and 3 / 4 { 1 ¯     0     1     0 } α which is consistent with the typical diffraction pattern of metastable β′ precipitates ( a = 2 × a α-Mg ≈ 0.64 nm, b = 8 × d ( 1 ¯     0     1     0 ) α -Mg ≈ 2.22     nm , c = c α-Mg ≈ 0.52 nm) with three variants along [0 0 0 1] α zone reported in the previous studies [6–8] about Mg–Gd–Y alloys. Actually, the SAED pattern overlaps diffraction patterns of three symmetrical orientation modes of β′ precipitates in [0 0 0 1] α zone. Fig. 5 shows the HRTEM image of oval plate-shape β′ phase observed along [0 0 0 1] α direction and corresponding Fourier transform pattern. The Fourier transform pattern is similar with the diffraction spots of one orientation modes of β′ precipitate which is shown in Fig. 4 (b). The HRTEM image indicated that the β′ precipitate has long-period ordered structure and is fully coherent with the matrix. Bright field image observed along [ 1 ¯     1     0     0 ] α direction from the sample near tensile fracture surface at 250 °C is shown in Fig. 6 . The β′ phases were observed distinctly with dense faults existed in it, and the slip line has not been found in the alloy. 3.2 Mechanical properties Fig. 7 shows the temperature dependence of ultimate tensile strength (UTS), tensile yield strength (TYS) and elongation of studied alloy. With increasing temperature, the tensile strength (the maximum UTS was reached at 250 °C, namely 345.5 MPa) and elongation increased, while the yield strength decreased slightly. When rods tested at 300 °C, the UTS and TYS decreased along with the increase of elongation. 3.3 Fractography Fig. 8 shows the SEM images of the tensile fracture surfaces at 20 °C and 250 °C. As shown in Fig. 8 (a), the SEM image of experimental alloy has shown the predominant intergranular fracture accompanied by minor transgranular rupture at 20 °C. The intergranular fracture facets were generally characterized by sharp ledges, which implied the mechanism affected by grain boundary slip. This fracture mode seems unbeneficial to the ductility of alloy. As for tensile fracture surface at 250 °C ( Fig. 8 (b)), visible plastic deformation was discovered on the intergranular fracture surface, and the elongation increased distinctly at the same time. It should be the activation of non-basal slip systems at elevated temperature that facilitated the grain deformability. 4 Discussion In present experiences, the aged Mg–12Gd–2Y–0.5Sm–0.5Sb–0.5Zr alloy exhibited excellent tensile strength at ambient and elevated temperature, and the mechanical properties have been influenced by the multi-strengthening mechanism as shown below. The solid solution strengthening of alloys arises from the elastic interaction between the strain field of a solute and that of dislocation. Not only the atomic size and/or shear modulus misfits, but also the electron state between the atoms should be considered to understand the effect of solute additions [9] . The high solid solution of the RE elements plays a positive role in the superior mechanical properties. The RE element atoms replace Mg atoms to form a random substitutional solid solution. At about 200 °C, the Gd will dissolve up to 3.8 wt.%, and Y will dissolve up to 2.2 wt.%, and Sm will dissolve up to 0.4 wt.% [10] . The RE elements atoms are bigger than the RE atoms, and, in squeezing into the Mg structure, generate stresses. These stresses roughen the slip plane, making it harder for dislocations to move, and thereby increases the dislocation yield strength. In general, the addition of high solid solution RE element could greatly enhance the tensile properties [11] . The superior mechanical properties could attribute to the grain-boundary strengthening. It is because that a dislocation passing into the near grains or dendrites of different orientations has to change its direction of motion, and the atomic disorder within a boundary region will result in a discontinuity of slip planes from one to the other. Furthermore, the existence of granular Mg–RE compounds could effectively hinder basal plane slip, and change the orientation of precipitates to inhibit the matrix deformation, especially at elevated temperatures. Suitable quantities of Mg–RE phase distributed along gain boundaries and interior grains, it can improve the strength of the alloy matrix, but excessive Mg–RE phase easy to initiate the cracks under tensile test, and the interior cracks are easy to connect and form the main crack, which accelerates the alloy failure [12] . For experimental alloy, the dispersed β′ precipitates interior the grain strengthened the Mg matrix, and can effectively prevent the grain deformation and dislocation glide. The oval plate-shape β′ precipitates, which formed on prismatic planes of Mg matrix in dense triangular arrangement, were vertical to basal plane of Mg. At the same time, the β′ precipitate is formed in the T6 treatments, exhibits semi-coherent structure with Mg matrix and remains stable during the subsequent tensile test at high temperature experiment, which indicates that the β′ phases in Mg alloy have a good thermal stability and could effectively prevent sliding of basal plane [11] . So the β′ phase could provide the effective obstacles to basal dislocation slip even at 250 °C. Fig. 9 shows the stress–strain curves of experimental alloy at 20 °C, 200 °C and 250 °C. As temperature increased to 250 °C from 20 °C, the yield strength decreased by 1.6%. The difference value between tensile strength and yield strength turned larger (i.e., from 19.6 MPa at 20 °C to 100.5 MPa at 250 °C). When temperature increased from 20 °C to 250 °C, the slight decrease of yield strength and significant reinforcement of deformation strengthening were the reasons that improve the tensile strength monotonously. At 300 °C, the tensile strength decreased. Commonly, the prismatic precipitates are thought as less effective obstacles to prevent non-basal slip and cross slip which activated at 250 °C or above, because they are not perpendicular to the slip planes now and could easily change the semi-coherent crystal lattice so that the effective area decrease as an obstacle to hinder dislocation slip [7] . We considered that the following factors help to comprehend the result. The magnesium has the hexagonal structure, and the basal and the non-basal slip systems operate simultaneously with increasing temperature [13] . At the same time, the β′ phase has LPO structure. It is reported [14] that in ordered structure the dislocations must travel in pairs—a leading dislocation and a trailing dislocation, and the dislocations in pairs can hardly climb or cross-slip. Therefore, the high-temperature strength can be increased significantly. Several materials, including beryllium, exhibit an anomalous flow stress increase at increasing temperature. Caillard and Martin [15] believed that the increasing of the recombination energy R (the recombination energy per unit length of dislocation) caused the difficulty of cross slipping, and induced the strength anomaly as the temperature rises. All cross slip mechanisms are very sensitive to the recombination energy R , i.e., the stacking fault energy. In the experiment, dense faults in LPO structure β′ phases ( Fig. 6 ) are detected in the alloy at 250 °C. Based on above discussions, it is reasonable to believe that the hard [16] , ductile and thermally stable LPO β′ phases provide important strengthening sources in the alloy, especially at elevated temperatures. In summary, it is rational to believe that the solid solution strengthening of RE, the precipitates strengthening and the LPO strengthening of β′ phases, together with grain-boundary strengthening and secondary strengthening, are all strengthening sources for experimental alloy. Because of all these strengthening sources, the alloy in T6 state exhibits high strength at both ambient and elevated temperatures, up to 300 °C. For anomalous temperature dependence of tensile strength, with the result that they will not restrict dislocation motion efficiently when temperature is higher than 0.5 T m , the mobility of solution atoms will be much higher than that of dislocations [15] . So it should be due to the effective obstruction of LPO β′ precipitates to dislocations in elevated temperature. However, there may be other factors that have not been considered in this discussion. The increasing temperature decreases the critical resolved shear stress of slip systems. The existence of RE solute atoms affects the dislocation substructures, especially, the non-basal slip of the dislocations. Moreover, the alloying can affect the materials constants; such as stacking fault energy, shear modulus and diffusion coefficient, which are important parameters in determining the mechanical properties [17] . Detailed experimental and analysis (for example, the contribution of each strengthening mechanism at low and high temperatures) need further research. 5 Conclusions (1) For aged Mg–12Gd–2Y–0.5Sm–0.5Sb–0.5Zr alloy, with temperature increase from 20 °C to 300 °C, the UTS increased and reached the maximum at 250 °C, namely 345.5 MPa. The elongation increased monotonously and reached the maximum value at 300 °C. The UYS decreased slightly. (2) The remarkable high strength of experimental alloy was mainly associated with solution strengthen of RE and precipitation strengthening of dispersive LPO structure β′ precipitates in Mg matrix. (3) The LPO β′ precipitates provide important strengthening sources in experimental alloy, especially at elevated temperatures. Acknowledgements This research was supported by Henan Research Program of Application Foundation and Advanced Technology ( 102300410018 ), Allotment Planning for College Excellence Young Teachers in Zhejiang Province , Key Research Program of Taizhou Vocational & Technical College ( 2011ZD07 ) and Zhejiang Industrial Technology Project for Public Welfare. References [1] J. Wang J. Meng D.P. Zhang D.X. Tang Mater. Sci. Eng. A 156 2007 78 84 [2] T. Honma T. Ohkubo S. Kamado Acta Mater. 55 2007 4137 4150 [3] Q.M. Peng X.L. Hou L.D. Wang Mater. Des. 30 2009 292 296 [4] Y. Gao Q.D. Wang J.H. Gu Jinhai J. Rare Earth 26 2008 298 302 [5] Z. Yang J.P. Li Y.C. Guo T. Liu F. Xia Z.W. Zeng M.X. Liang Mater. Sci. Eng. A 454 2007 274 280 [6] T. Honma T. Ohkubo K. Hono S. Kamado Mater. Sci. Eng. A 395 2005 301 306 [7] S.M. He X.Q. Zeng L.M. Peng X. Gao J.F. Nie W.J. Ding J. Alloys Compd. 427 2007 316 323 [8] X. Gao S.M. He X.Q. Zeng L.M. Peng W.J. Ding J.F. Nie Mater. Sci. Eng. A 431 2006 322 327 [9] L. Gao R.S. Chen E.H. Han J. Alloys Compd. 481 2009 379 384 [10] K.-J. Li Q.-A. Li X.-T. Jing J. Chen X.-Y. Zhang Q. Zhang Scripta Mater. 60 2009 1101 1104 [11] Q. Peng X. Hou L. Wang Y. Wu Z. Cao L. Wang Mater. Des. 30 2009 292 296 [12] Q. Peng L. Wang Y. Wu L. Wang J. Alloys Compd. 469 2009 587 592 [13] M. Suzuki H. Sato K. Maruyama H. Oikawa Mater. Sci. Eng. A 252 1998 248 255 [14] D.D. Yin Q.D. Wang Y. Gao C.J. Chen J. Zheng J. Alloys Compd. 509 2011 1696 1704 [15] D. Caillard J.L. Martin J. Phys. France 50 1989 2455 2473 [16] J.T. Guo The Basic Theory Application of High-Temperature Alloy Materials Science 2008 Science Press Beijing pp. 26–48 [17] L. Gao R.S. Chen E.H. Han J. Alloys Compd. 472 2009 234 240
What problem does this paper attempt to address?