Effect of heat treatment on tension–compression yield asymmetry of AZ80 magnesium alloy
Chengling Lv,Tianmo Liu,Dejun Liu,Shan Jiang,Wen Zeng
DOI: https://doi.org/10.1016/j.matdes.2011.04.060
IF: 9.417
2012-01-01
Materials & Design
Abstract:The tension–compression yield asymmetry of extruded AZ80 alloy under different heat treatments was studied using uniaxial tensile and compressive tests at room temperature. It was found that aging treatment was much more effective in reducing the asymmetry than solution treatment. Finer twins and lower twin volume fraction in aged alloy were observed by optical microscope and X-ray diffraction. In order to understand the tiny effect of solution treatment on the asymmetry, quadrangular prism-shaped atomic group unit model was used to show that alloying elements in solution suppressed the {1 0 −1 2} twinning. The smaller grain size also inhibited twinning during deformation. On the other hand, the macroscopical interaction between the precipitation and twin was proposed to illustrate the significant impact of aging treatment on the asymmetry. Keywords A. Mg metal matrix C. Heat treatments G. X-ray analysis 1 Introduction Magnesium alloys have a potential future in many applications due to their excellent physical properties such as high specific strength, high stiffness, and low density [1–3] . However, in contrast to cubic metals the yield behavior of most magnesium alloys differs significantly between tension and compression [4,5] , thereby preventing them from widespread application. The ratio between tensile and compressive yield strength (denoted the yield asymmetry ratio) is used to describe the extent of tension–compression yield asymmetry. As engineering materials, how tension–compression yield asymmetry of Mg alloys can be reduced so as to suit the multiple applications has been one of the central themes. It is well known that Mg and its alloys are low-symmetry materials with hexagonal-close packed (hcp) structures that have axial ratios (c/a) of around 1.633 [6] . When the axial ratio is less than 3 normally only the {1 0 −1 2} 〈1 0 −1 1〉 twinning is activated by a compressive stress parallel to the basal plane or a tensile stress perpendicular to the basal plane, which means that {1 0 −1 2} twinning is not favored in tension whereas it is preferred in compression along the extrusion axis. Thus, the yield asymmetry can be attributed to the prevalence of {1 0 −1 2} twinning [7,8] . Earlier studies in tension–compression yield asymmetry of Mg alloys can be summarized mainly in five aspects. Firstly, this yield asymmetry can be alleviated by weakening texture [9] . For example, the yield asymmetry is lower in cast alloys than wrought alloys [10] . Secondly, aluminum content affects the yield asymmetry. Bohlen et al. [11] have shown that the difference between tensile and compressive yield strength generally decreases with increasing content of aluminum in the alloy, referring to the same grain size. Thirdly, Chang et al. [8] and Bohlen et al. [11] have observed that the yield asymmetry also decreases with decreasing grain size. Fourthly, the effect of load direction on twinning activity and the asymmetry in an extruded Mg–3Al–1Zn alloy has been studied by Yin et al. [12] . When load angle was 45°, the yield asymmetry ratio equaled 1.02. Finally, the mechanical properties and yield behaviors of the Mg alloys can be tuned severely through heat treatment [13–15] . In fact, this relevance has already been highlighted by the claim that aging treatment can reduce the yield asymmetry [16] , because the presence of precipitates modifies the process of twinning [17,18] . Gharghouri et al. [17] studied a Mg–7.7 at.% alloy and proposed that the difficulty of twin boundaries migration in propagating through a high density of precipitates resulted in the interaction between twinning and precipitates. After that, Stanford et al. [18] reported reductions in twin size and total twin volume fraction in an aged Mg–5% Zn alloy. Although much effort has been put into the investigation of yield behaviors, the yield asymmetry of Mg alloys in solution and aging heat treatments have not yet been fully clarified. The current study is then designed to investigate the effect of different heat treatments on twinning during deformation and therefore tension–compression asymmetry of AZ80 Mg alloy at room temperature. 2 Experimental procedure The starting material for this study was an extruded rod of AZ80 (Mg–8 wt.% Al–0.5 wt.% Zn) alloy. The extruded specimens were first solution treated at 410 °C for 12 h followed by water quenching, and then aged at 230 °C for 8 h. To explore the role that solution and age treatment play in modifying the tension–compression asymmetry, tension (5 mm diameter and 25 mm gauge length) and compression (10 mm diameter and 12 mm height) specimens were machined from the extruded and heat treated materials. Mechanical tests were then carried out on a CMT5105 material test machine at a constant strain rate of 10 −3 s −1 . The yield strength was measured as 0.2% proof stress σ 0.2 . The microstructure and morphology of the specimens was characterized by optical microscopy (OM) and scanning electron microscope (SEM) using an accelerating to voltage of 20 kV. X-ray diffraction (XRD) was performed using a D/Max-1200X diffractometry with Cu Kα radiation operated at 30 kV and 100 mA, in order to identify the crystallographic orientation. 3 Results 3.1 Effect of heat treatment on tension–compression yield asymmetry The tension–compression yield asymmetry described above is clearly illustrated in the tensile and compressive engineering stress–strain curves of as-extruded and solution-treated samples, as shown in Fig. 1 a and b. It can be seen that both the yield strengths and the rate of work hardening are lower in compression than in tension. The yield asymmetry ratios are 0.65 and 0.67 in as-extruded and solution samples, respectively. No marked difference between the curves of these two states suggests that solution treatment has a slightly impact on decreasing the asymmetry. Shown in Fig. 1 c are the engineering tensile and compressive stress–strain curves for the aged alloy. In this state the stress–strain curves show a phenomenon of precipitation hardening and the ductility in tension is lower than in compression. The asymmetry ratio is 0.79, higher than the as-extruded and solution-treated samples. Therefore, it is indicated that aging treatment is much more effective in reducing the tension–compression yield asymmetry for the extruded AZ80 alloy compared with solution treatment. 3.2 Microstructure and XRD observations The microstructures of the as-extruded and heat treated AZ80 alloy are shown in Fig. 2 . The OM and SEM micrograph of as-extruded sample ( Fig. 2 a) shows the small dark second phase recognized at grain boundaries with the average grain size of 22 μm. The solution heat treatment (410 °C for 12 h) results in almost complete dissolution of β -Mg 17 Al 12 in the matrix and provides an equiaxed microstructure, as shown in Fig. 2 b. However, precipitations can be seen at grain boundaries and within grains by aging treatment in Fig. 2 c. Moreover, Fig. 2 d reveals the magnified view of precipitates in aged condition, which consists of two types, continuous (general precipitates) and discontinuous (cellular precipitates) [19,20] . The average initial grain sizes for both the solution-treated and aged samples are around 25 μm. To gain insight into the texture evolution of the material, the X-ray diffraction (XRD) is employed. Fig. 3 a shows the crystallographic orientation of as-extruded sample. The intensity of the (0 0 0 2) peak, (denoted as I (1010) / I (0002) ratio [21] ), is ∼0.19, so the initial texture can be roughly considered as having a ring fiber texture with the basal planes parallel to the extrusion direction. This initial texture does not allow any twinning under tensile loading but maximum twinning under compressive loading. Therefore, this starting texture could activate the {1 0 −1 2} twinning at a low degree of deformation. In order to evaluate whether the differences between solution and aged samples can be attributed to the differences in twinning behavior, the crystallographic orientation after ∼3% compressive deformation is measured as shown in Fig. 3 b and c. It can be seen that there is a marked difference in I (1010) / I (0002) ratio, thus the (0 0 0 2) intensity, between the two specimens. The aged specimen shows a much weaker the (0 0 0 2) intensity compared to solution-treated condition. Since {1 0 −1 2} twinning leads to a reorientation of 86.3° of the crystal lattice, all the basal planes in twinned lattices lie nearly perpendicular to the extrusion direction, that is to say, parallel to (1 0 −1 0) after compressive plastic deformation. According to the sharp texture in the starting material ( Fig. 3 a), with virtually no grains lying near (0 0 0 1), it can be assumed that the large reorientation originates from {1 0 −1 2} twinning. These XRD data can thus be used to estimate the volume fraction of twins in the microstructure. For the same strain level of 3% compression, the I (1010) / I (0002) ratio in solution treated specimen is ∼0.56, while the I (1010) /I (0002) ratio in aged specimen is ∼0.37. Hence, it is reasonable to conclude that the aged specimen has a lower total volume fraction of {1 0 −1 2} twins than solution-treated specimen. Further evidence of lower tensile twinning activity in the aged sample can be obtained from microstructural observations. Fig. 4 shows the microstructures in different states after having been strained 3% in compression along extrusion direction. It can be seen that the twins in the presence of precipitates are finer and smaller ( Fig. 4 b) compared with the solution-treated sample ( Fig. 4 a). Meanwhile, the macroscopical twin boundary–precipitate interactions is seen including cases of twins bowing around, bypassing and impinging precipitates shown in Fig. 4 b. Qualitatively, it is observed that the fraction of tensile twins in the aged material is lower than in the solution-treated case, in agreement with the analysis of XRD. These results indicate that the effect of heat treatment on tension–compression yield asymmetry is mainly attributed to {1 0 −1 2} tension twinning. 4 Discussion 4.1 Effect of alloying elements on {1 0 −1 2} twinning It has been suggested [22,23] that the presence of alloying elements in solution can alter the axial ratio, and therefore affect the balance of slip deformation modes. As mentioned previously, the chosen alloy contains significant amounts of Al and Zn. The small amounts of precipitates at grain boundary are dissolved into matrix by solution treatment, thereby changing the axial ratio. To understand how alloying elements influent the {1 0 −1 2} twinning and yield asymmetry, the relationship between the axial ratio ( γ ) and {1 0 −1 2} twinning is provided based on our earlier work [24] . Fig. 5 presents the quadrangular prism-shaped atomic group (QPAG) unit model for {1 0 −1 2} twinning in Mg alloys. From this figure, it can be seen that the atomic groups labeled DCGH and ABF are eligible to be QPAG units because they can rotate by an angle and make a vector as a whole, but their relative atomic positions intact during twinning. In order to calculate the rotational angle of the QPAG angle simply, the ABF unit is chosen. In this sense, the relationship between the rotational angle ( α ) and axial ratio ( γ ) can be expressed by: (1) α = arctan Δ z B | y B - y A | - Δ y B = arctan 2 sin 2 arctan 3 γ 3 - 3 γ 3 - 2 cos 2 arctan 3 γ 3 where Δ z B , Δ y A and Δ y B represent vector components of atom A and B along the y and z axis. In light of Eq. (1) and keeping in mind that the axial ratios examined by XRD in the extrusion and solution states are 1.6226 and 1.625, respectively, then the rotational angle can be calculated as follows: (2) α solution - treated ≈ 15.91 ∘ (3) α as - extruded ≈ 15.85 ∘ Evidently, the QPAG unit in solution treated condition rotates a higher angle than that in extrusion condition without considering the grain size effect (Eqs. (2) and (3) ). Since the atomic motion in this model can be simply ascribed to the rotational motion of QPAG units, the degree of the difficulty for {1 0 −1 2} twinning in Mg alloys is determined by the magnitude of the relative motion of these units. This indicates that solute alloying elements (Al and Zn) can inhibit the {1 0 −1 2} twinning, and are therefore likely to affect the critical resolved shear stresses of basal and non-basal slip. Akhtar and Teghtsoonian [25] have reported that single crystals of Mg and a series of alloys containing Al and Zn solute can induce first order {1 0 −1 0}〈1 1 −2 0〉 prismatic slip while suppress both basal slip and {1 0 −1 2} twinning, which is agree with our calculation. 4.2 Effect of grain size on {1 0 −1 2} twinning The yield strength ( σ ) can be expressed in terms of the Hall–Petch law as follows: (4) σ = σ 0 + kd - 1 / 2 where σ 0 and k are the Hall–Petch’s friction stress and stress intensity factor, respectively. In general, the yield strength decreases with increasing grain size ( d ). This can be attributed to the fact that massive dislocation slip as well as twinning does occur with lower activity. When the grain size turns smaller, basal and non-basal slip and even grain boundary sliding mechanisms can be active, thereby suppressing twinning during deformation, in other words, twinning is more likely to happen in bigger grain. This is in accord with Jan Bohlen et al.’s report that the yield asymmetry decreases with decreasing grain size [11] . In our results, the bigger grain size ( Fig. 2 a and b) in solution-treated sample can facilitate {1 0 −1 2} twinning and therefore worsen yield asymmetry, while the solute alloying elements show opposite trend. This accounts for the tiny impact of solution treatment on reducing the asymmetry of AZ80 alloy. 4.3 Precipitate twin interactions The XRD and optical microscopy data presented in the results section show that extensive {1 0 −1 2} twinning occurs during compression, but the twins are smaller and have a lower volume fraction in the presence of precipitates material. Jain and Poole [16] have also been reported the familiar results by EBSD maps. These observations suggest that the reduction in the tension–compression asymmetry is attributed to reduced rates of twinning in the presence of precipitates. Although the effect of the precipitates on the nucleation of twinning is unclear, the twin growth is limited. Clearly, how precipitates inhibit twin growth needed to be further understood. In a particle-free matrix the {1 0 −1 2} twin can easily grow laterally. In the case of an aged alloy, if the lateral growth is limited, the grain then is subject to higher stress because the imposed strain cannot be accommodated by the twin. Since the particle is not sheared during the twinning [17,18] , as the twinning ledge approaches the particle there will be some back-stress resulting from the rigid particle inhibiting the free shear of the moving twin boundary. Based on many examples of twin boundary–precipitate interactions, including cases of twins bowing around and bypassing precipitates [17] and presented microstructures in Fig. 4 b, the macroscopical interaction between precipitation and twin can be roughly grouped into three categories as shown in Fig. 6 : (i) precipitates are beside the twin boundary ( Fig. 6 a). The back-stress can prevent one side of {1 0 −1 2} twin boundary moving laterally and make the other side move in the opposite direction, which results in twins finer; (ii) precipitates are face-to-face with the twin boundary ( Fig. 6 b). Under this condition, it is likely to restrain the length of the twin; and (iii) there is an obvious angle between precipitates and twin boundary ( Fig. 6 c). In this situation, both the two former cases can take place. The twin boundary become curved if the stress is enough. Whatever the situation may be, this back-stress must be overcome by the applied stress. When the shear takes place at the matrix–precipitate interface, the movement of atoms toward the precipitate must be accommodated by the surrounding material. It might be that the twin shape change in Fig. 6 c is mainly accommodated by basal slip inside the twin, rather than in the parent material because the precipitate also restricts the basal slip in the parent region. In addition, non-basal slip with 〈 a 〉 type might be present due to the stress concentrations near particle–matrix boundaries [26] . The results presented here suggest that Mg 17 Al 12 precipitates affect twin growth and deformation modes, which is attributed to reduce the yield asymmetry in AZ80 alloy. 5 Conclusions Solution (410 °C-12 h) and aging (410 °C-12 h + 230 °C-8 h) treatments have been applied to the extruded AZ80 Mg alloy, aimed at investigating the effect of heat treatment on twinning during deformation and therefore tension–compression yield asymmetry at room temperature. The following conclusions have been made: (1) Aging treatment is much more effective in reducing the tension–compression yield asymmetry than solution treatment for extruded AZ80 Mg alloy. The asymmetry ratio increases from 0.65 to 0.67 in extruded and solution-treated samples to 0.79 in aged sample. (2) XRD and optical microscopy analysis show that the aged alloy produces finer twins and lower twin volume fraction in response to 3% compression, compared to the precipitate-free specimen in solution-treated condition. (3) Quadrangular prism-shaped atomic group (QPAG) unit model for {1 0 −1 2} twinning is used to show that alloying elements (Al and Zn) in solution suppress the {1 0 −1 2} twinning. The bigger grain size in solution-treated sample, however, presents an opposite impact on twinning during deformation. These analysis may account for the tiny effect of solution treatment on tension–compression yield asymmetry. (4) The macroscopical interaction between precipitation and twin is proposed to illustrate the significant impact of aging treatment on reducing the asymmetry. Acknowledgements This work was supported in part by the Fundamental Funds for the Central Universities (CDJXS10131154) and the Research National 973 Major Project of China, “The Key Fundamental Problem of Processing and Preparation for High Performance Magnesium Alloy”, under Grant No. 2007CB613700. References [1] H. Takuda T. Enami K. Kubota N. Hatta The formability of a thin sheet of Mg–8.5Li–1Zn alloy J Mater Process Technol 101 2000 281 286 [2] J.A. Del Valle M.T. Perez-Prado O.A. Ruano Texture evolution during large-strain hot rolling of the Mg AZ61 alloy Mater Sci Eng A 355 2003 68 78 [3] J. Bohlen S.B. Yi J. Swiostek D. Letzig H.G. Brokmeier K.U. Kainer Microstructure and texture development during hydrostatic extrusion of magnesium alloy AZ31 Scripta Mater 53 2005 259 264 [4] S. Kleiner P.J. Uggowitzer Mechanical anisotropy of extruded Mg–6% Al–1% Zn alloy Mater Sci Eng A 379 2004 258 263 [5] E.A. Ball P.B. Prangnell Tensile-compressive yield asymmetries in high strength wrought magnesium alloys Scripta Mater 31 1994 111 116 [6] S.X. Song J.A. Horton N.J. Kim T.G. Nieha Deformation behavior of a twin-roll-cast Mg–6Zn–0.5Mn–0.3Cu–0.02Zr alloy at intermediate temperatures Scripta Mater 56 2007 393 395 [7] S.R. Agnew M.H. Yoo C.N. Tome Application of texture simulation to understanding mechanical behavior of Mg and solid solution alloys containing Li or Y Acta Mater 49 2001 4277 4289 [8] L.L. Chang Y.N. Wang X. Zhao M. Qi Grain size and texture effect on compression behavior of hot-extruded Mg–3Al–1Zn alloys at room temperature Mater Charact 60 2009 991 994 [9] S.M. Yin C.H. Wang Y.D. Diao S.D. Wu S.X. Li Influence of grain size and texture on the yield asymmetry of Mg–3Al–1Zn alloy J Mater Sci Technol 27 2011 29 34 [10] M.R. Barnett C.H.J. Davies X. Ma An analytical constitutive law for twinning dominated flow in magnesium Scripta Mater 52 2005 627 632 [11] J. Bohlen P. Dobroň J. Swiostek D. Letzig F. Chmelík P. Lukáč On the influence of the grain size and solute content on the AE response of magnesium alloys tested in tension and compression Mater Sci Eng A 462 2007 302 306 [12] D.L. Yin J.T. Wang J.Q. Liu X. Zhao On tension–compression yield asymmetry in an extruded Mg–3Al–1Zn alloy J Alloys Compd 478 2009 789 795 [13] Q. Peng X. Hou L. Wang Y. Wu Z. Cao L. Wang Microstructure and mechanical properties of high performance Mg–Gd based alloys Mater Des 30 2009 292 296 [14] D.K. Xu L. Liu Y.B. Xu E.H. Han The effect of precipitates on the mechanical properties of ZK60-Y alloy Mater Sci Eng A 420 2006 322 332 [15] S. Liang D. Guan L. Chen Z. Gao H. Tang X. Tong Precipitation and its effect on age-hardening behavior of as-cast Mg–Gd–Y alloy Mater Des 32 2010 361 364 [16] J. Jain W.J. Poole C.W. Sinclair M.A. Gharghouri Reducing the tension–compression yield asymmetry in a Mg–8Al–0.5Zn alloy via precipitation Scripta Mater 62 2010 301 304 [17] M.A. Gharghouri G.C. Weatherly J.D. Embury The interaction of twins and precipitates in a Mg–7.7 at.% Al alloy Philos Mag A 78 1998 1137 1149 [18] N. Stanford M.R. Barnett Effect of particles on the formation of deformation twins in a magnesium-based alloy Mater Sci Eng A 516 2009 226 234 [19] J.B. Clark Age hardening in a Mg-9 wt.% Al alloy Acta Mater 16 1986 141 152 [20] S. Celotto TEM study of continuous precipitation in Mg–9 wt% Al–1 wt% Zn alloy Acta Mater 48 2000 1775 1787 [21] Y.N. Wang J.C. Huang The role of twinning and untwinning in yielding behavior in hot-extruded Mg–Al–Zn alloy Acta Mater 55 2007 897 905 [22] M. Suzuki H. Sato K. Maruyama H. Oilawa Creep behavior and deformation microstructures of Mg–Y alloys at 550 K Mater Sci Eng A 252 1998 248 255 [23] S. Ando H. Tonda Non-basal slip in magnesium–lithium alloy single crystals Mater Trans JIM 41 2000 1188 1191 [24] S. Jiang T.M. Liu L.W. Lu W. Zeng Z.C. Wang Atomic motion in Mg–3Al–1Zn during twinning deformation Scripta Mater 62 2010 556 559 [25] A. Akhtar E. Teghtsoonian Solid solution strengthening of magnesium single crystals—II the effect of solute on the ease of prismatic slip Acta Mater 17 1969 1351 1356 [26] M.A. Gharghouri G.C. Weatherly J.D. Embury J. Root Study of the mechanical properties of Mg–7.7 at% Al by in-situ neutron diffraction Philos Mag A 79 1999 1671 1695