This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
, q>2} can be expressed into the rotation matrix given by Equation 9. There are 12 misorientations Ag,, between two orientations g, andg, where g is the orientation of an integration point in the grain g and S is the mean orientation of all the orientations in the same grain. Agj can be calculated by Equation 10 where Mj is the symmetry operator which depends on the symmetry of the HCP crystal system. Usually, the minimum rotation angle A0j defines the misorientation angle or disorientation angle which is expressed as Equation 11. ' cos^cospj-siiKpiSin^cos® cosç>, s i n ^ ; - s i n ^ | COS?J3COS(I>
sinp, c o s p ; + c o s ^ s i n ^ . c o s O
sinç>2sin
- s i n ^ sirnp-.+cos^, cosç>: cosŒ
cos^,sin
-COSÇ0, sin(t>
cosO
siniflsinO
Figure 4: Pole figures for an AM30 magnesium alloy showing (a) the experimental initial texture and (b) the experimental channel die compression at a strain of 30% and (c) the finite element crystal plasticity simulation result at a strain of 30%.
(9) As, = « , - s , - « '
:
'4?„
4S,2
A?2,
AS22
(10) Afe
A&J,
1 1 AÖ, = (arccos( A g n + A f e - - ^ ? - )v)min/mwm
(11)
585
important role on orientation spread, stress and strain localization and intergranular heterogeneous plasticity. Conclusions Plane strain compression simulations of a three dimensional microstructure of HCP magnesium AM30 were performed using an elastic-plastic crystal plasticity model and the finite element method. The mechanical response, global texture evolution and the intergranular heterogeneous plasticity of discrete grain were captured by the 3D Voronoi microstructure channel die compression simulation. Simulation results showed that the effect of grain interaction could play an important role on global texture evolution, the orientation spread and the local heterogeneity deformation of one grain. Acknowledgements The authors are grateful to the financial support from the Department of Energy, Contract No. DE-FC-26-06NT42755, and the Center for Advanced Vehicular Systems (CAVS) at Mississippi State University.
Figure 5: For an AM30 magnesium alloy (a) shows mesh of the initial grain number 123, (b) the local Mises stress and (c) local strain 633 distribution of grain number 123 at a strain of 30%, (d) the orientation distribution of the grain number 123 at a strain of 30%. The cross symbols represent the orientations, the circle symbol represents the initial orientation of grain number 123 and the cube symbol represents the mean orientation among all the orientations, and (e) the orientation spread of the grain number 123 at a strain of 30%.
References D. Raabe, "Yield surface simulation for partially recrystallized aluminum polycrystals on the basis of spatially discrete data," Computational Materials Science, 19 (2000), 13-26. A. Fjeldly and H.J. Roven, "Observations and calculations on mechanical anisotropy and plastic flow of an AlZnMg extrusion," Acta Materialia, 44(1996), 3497-3504. G Sachs, "Zur ableitung einer fleissbedingun," Zeichschrifi Verein Deutscher Ingenieur, 72 (1928), 734-736. GI. Taylor, "Plastic strain in metals," Journal of the Institute of Metals, 62 (1938), 307-324. D. Raabe, Z. Zhao, and W. Mao, "On the dependence of ingrain subdivision and deformation texture of aluminum on grain interaction," Acta Materialia, 50 (2002), 4379-4394. P. Van Houtte et al., "Deformation texture prediction: From the Taylor model to the advanced Lamel model," Internationaljournal of Plasticity, 21 (2005), 589-624. U.F. Kocks and H. Chandra, "Slip geometry in partially constrained deformation," Acta Metallurgica, 30 (1982), 695-709. W. Mao and Y. Yu, "Effect of elastic reaction stress on plastic behaviors of grains in polycrystalline aggregate during tensile deformation,"Atoen'a/.y Science and Engineering, A367 (2004) 277-281. R.A. Lebensohn and C.N. Tome, "A self-consistent anisotropic approach for the simulation of plastic deformation and texture development of polycrystals: Application to zirconium alloys," Acta Materialia, 41 (1993), 2611-2644. 10. D. Peirce, R.J. Asaro, and A. Needleman, "Material rate dependence and localized deformation in crystalline solids," Acta Metallurgica, 31 (1983), 1951-1976. 11. D. Peirce, R.J. Asaro, and A. Needleman, "Analysis of nonuniform and localized deformation in ductile single crystals," Acta Metallurgica, 30(1982), 1087-1119. 12. A. Alankar, Ioannis N. Mastorakos, and D.P. Field, "A
Figure 6: The finite element simulation stress-strain curve based on crystal plasticity model and the experimental stress-strain curve under channel die compression for AM30 magnesium alloy at 200°C.
CPFEM predicted textures of the 3D Voronoi microstructure that are consistent with experimental results as shown in Figure 4b and 4c. As shown by this figure, the measured and the predicted plane strain compression AM30 textures exhibit similar texture configuration. However, there is discrepancy that could stem from the fact that dynamic recrystallization, which typically happens in Mg alloys, is not accounted for in the simulations. Also, the CPFEM simulation stress-strain curve slightly overestimates the stress of the channel die compression. The reason may be due to the effect of grain interaction in the 3D Voronoi microstructure. Figure 5b and 5c present the Mises and e33 strain distribution in grain number 123 at strain 30%. The inhomogeneous deformation of grain number 123 is clearly proved by the observed stress and strain heterogeneous distribution. In addition, as shown in Figure 5d, the grain orientation distribution in grain number 123 displays the different rotation tendency along RD direction. As such, the orientation spread of grain number 123 presents a big scatter with two peaks, as shown in Figure 5e. Accordingly, the effect of grain interaction among this 3D Voronoi polycrystal should play an
586
13.
14.
15.
16. 17. 18. 19. 20. 21. 22.
23. 24. 25.
26.
dislocation-density-based 3D crystal plasticity model for pure aluminum," Ada Materialia, 57 (2009), 5936-5946. E.B. Marin and P.R. Dawson, "Elastoplastic finite element analyses of metal deformations using polycrystal constitutive models "Computer Methods in Applied Mechanics and Engineering, 165 (1998), 23-41. F. Roters et al., "Overview of constitutive laws, kinematics, homogenization and multiscale methods in crystal plasticity finite-element modeling: Theory, experiments, applications,"/fcta Materialia, 58 (2010), 1152-1211. Z. Zhao et al., "Influence of in-grain mesh resolution on the prediction of deformation textures in fee polycrystals by crystal plasticity FEM," Ada Materialia, 55 (2007), 23612373. G.B. Sarma and P.R. Dawson, "Effects of interactions among crystals on the inhomogeneous deformations of polycrystals ," Ada Materialia, 44(1996), 1937-1953. D.P. Mika and P.R. Dawson, "Effects of grain interaction on deformation in polycrystals ," Materials Science and Engineering, A257 (1998), 62-76. P.R. Dawson, D.P. Mika, and N.R. Barton, "Finite element modeling of lattice misorientations in aluminum polycrystals," ScriptaMaterialia, 47 (2002), 713-717. D.P. Mika and P.R. Dawson, "Polycrystal plasticity modeling of intracrystalline boundary textures," Ada Materialia, 47(1999), 1355-1369. E.B. Marin, "On the formulation of a crystal plasticity model" (Report SAND2006-4170, Sandia National Laboratories, CA, 2006). S. Groh et al., "Multiscale modeling of the plasticity in an aluminum single crystal "International Journal ofPlasticity, 25(2009), 1456-1473. E.B. Marin and P.R. Dawson, "On modeling the elastovisoplastic response of metals using polycrystal plasticity," Computer Methods in Applied Mechanics and Engineering, 165 (1998), 1-21. R. Hielscher and H. Schaeben, "A novel pole figure inversion method: specification of the MTEX algorithm," Journal ofApplied Crystallography, 41 (2008), 1024-1037. http://neper.sourceforge.net M. Blicharski, R. Becker, and H. Hu, "Deformation texture of channel-die deformed aluminum bicrystals with S orientations," Ada Metallurigica et Materialia, 41 (1993), 2007-2016. Y.L. Liu, H. Hu, N. Hansen, "Deformation and recrystallization of a channel die compressed aluminium bicrystal with (112)[111]/(123)[412] orientation," Ada Metallurigica et Materialia, 43 (1995), 2395-2405.
Magnesium Technology 2011 Edited by: Wim H. Sillekens, Sean R. Agnew, Neale R. Neelameggham, andSuveen N. Mathaudhu TMS (The Minerals, Metals & Materials Society), 2011
INVESTIGATION OF MICROHARDNESS AND MICROSTRUCTURE OF AZ31 ALLOY AFTER HIGH PRESSURE TORSION Jitka Vrâtnâ1, Milos Janeöek1, Josef Strâsky1, Hyoung Seop Kim2, Eun-Yoo Yoon2 'Charles University, Department of Physics of Materials, Ke Karlovu 5, CZ-12116 Czech Republic department of Materials Science and Engineering, POSTECH, Pohang, 790-784 Korea Keywords: AZ31 alloy, High pressure torsion, Microhardness evolution versatility and the scalability. Attractive results have been achieved using ECAP e.g. with aluminium alloys [8, 9], nevertheless the ECAP was less effective if applied to magnesium, namely dilute Mg alloys or pure Mg [10]. Recently, Horita [11] introduced a combined two-step processing route involving an initial extrusion step and subsequent processing by ECAP. This process designated by the acronym EX-ECAP, was used successfully to achieve UFG microstructures in many materials including magnesium alloys [12].
Abstract Cast commercial magnesium alloy AZ31 was processed by high pressure torsion (HPT) at room temperature for 1, 3, 5 and 15 rotations (strain ranged from 1 to 7). Microstructure evolution with strain imposed by HPT was observed by light and electron microscopy. HPT was shown to be a very effective method of grain refinement. The initial coarse grain structure was refined by a factor of almost 200 already after one HPT turn (s = 4). Mechanical properties were investigated by detailed 2-D microhardness measurements. HPT straining was found to introduce a radial inhomogeneity in the material which is manifested by a pronounced drop in the center and the maximum near the specimen periphery. With increasing strain due to HPT this inhomogeneity is continuously smeared out tending to saturate with increasing strain. Integrated 3-D meshes across the total surface of disks revealed the undulating character of microhardness variations. The strain imposed by HPT was shown to saturate with increasing number of HPT turns.
Although general principles of high pressure torsion were first proposed many years ago, it has become of general scientific interest only recently. It is only within approximately the last 5 years that numerous extensive reports documenting the processing and properties of materials fabricated by HPT have started appearing in the scientific literature showing that HPT is more effective technique of grain refinement than other SPD techniques. Extensive investigation of microstructure and mechanical properties of pure metals and solid solution alloys with FCC structure were conducted and reported in the scientific literature, e.g. in Cu [13-15], Ni [16] and Al [17-19]. Reports of successful application of HPT processing of metals with BCC structure can be found in the literature, e.g. on steels, Mo, Cr and W [20-21]. BCC metals were found to be more difficult to deform by HPT than FCC metals.
Introduction Due to high specific strength, magnesium alloys are very attractive and promising materials for structural components in automotive and aerospace industries. However, the use of Mg alloys in more complex applications is limited because of problems associated with low ductility, poor corrosion and creep resistance. The limited ductility is a consequence of the lack of independent slip systems and the large difference in the values of the critical resolved shear stress in the potential slip systems. Moreover, the occurrence of strong deformation textures and stress anisotropy in magnesium alloys reduces significantly the variety of possible industrial applications [1].
On the other hand, data about physical properties of UFG metals and alloys with hexagonal lattice are scarce. There are reports of HPT processing of titanium of commercial purity used for biomédical and dental applications [22] and zirconium for the use in surgical implants [23]. The reports describing the use of HPT with Mg alloys are almost lacking. Only recently Horita et al. reported significant grain refinement in Mg-9% Al alloy [24] and AZ61 [25].
The properties of magnesium alloys can be improved by refining the grain size to the submicrocrystalline (grain sizes in the range of 100-1000 nm) or even nanocrystalline level (grains sizes smaller than 100 nm) [2-3]. A variety of new techniques have been proposed for the production of the ultra-fine grain (UFG) structure in materials. All these techniques rely on the imposition of heavy straining and thus introduction very high density in the bulk solid material. Since these procedures introduce severe plastic deformation (SPD) into the material, it became convenient to describe all of these operations as SPD processing.
This work is therefore motivated by this fact and its main objective is to extend this missing knowledge of the properties of UFG Mg based alloys processed by HPT. The objective of the paper is to provide a detail analysis of microstructure evolution in AZ31 alloy subjected to HPT straining and to correlate it with the observation of microstructure. Experimental Commercial AZ31 alloy, with a nominal composition of Mg3%Al-l%Zn in the initial as cast condition was used in this investigation. Prior to high pressure torsion the alloy was homogenized at 390°C for 12 hours. After homogenization the disk specimens of the diameter of 20 mm and the thickness of 1 mm were cut from the billet. These specimens were processed by high pressure torsion at room temperature for 1, 3, 5 and 15 rotations by applying the hydrostatic pressure of 2.5 GPa. The experimental setup schematically illustrated at Fig. 1 comprises
Several processes of SPD are now available but only three of them receiving the most attention at present time, in particular equal-channel angular pressing (ECAP) [4], accumulative rollbonding (ARB) [5] and high-pressure torsion (HPT) [6]. Equal-channel angular pressing, first reported by Segal [7], became probably the most popular technique of SPD due to the combination of its effectiveness in producing UFG structure, the
589
two anvils. The upper anvil is fixed with a load cell mounted on its top allowing measuring the hydrostatic pressure which is applied to the specimen during straining. The lower anvil with a sample placed in the groove is first lifted to its final position pressing the specimen to the symmetrical groove in the upper anvil. Once the operating hydrostatic pressure is reached, specimen compression is stopped and maintained for 10 seconds. Then the torsional straining stage starts by rotation of the lover anvil with the constant speed of 1 rpm while maintaining the preset hydrostatic pressure.
Results The microstructure of AZ31 alloy in the initial condition (after homogenization treatment) is shown in Fig. 3. The microstructure consists of almost equiaxed grains with the average size of approximately 150 um.
Figure 3. Microstructure of AZ31 alloy in the initial condition. Microhardness profiles, i.e. the dependence of the microhardness on the distance r from the center of the disk were measured along two perpendicular lines passing through the center of the disc (see also Fig. 2). Fig. 4 displays the values of Vickers microhardness measured in linear traverses across the diameter of disks of AZ31 alloy after performing HPT at room temperature at different number of rotations (N=l, 3, 5 and 15). Each data point was obtained as the average of 4 Hv values obtained at the same distance from the center of the disk. The scatter of the data is in the range of approximately 10% (+/- 5 MPa). For clarity the error bars are not shown in Fig. 4. The base line of the compressed (N=0) specimen having the average value of 75 MPa is also shown in Fig. 4 to see the influence of compression and subsequent torsional straining on Hv values. Note that the average microhardness of the annealed specimen is 58 +/- 3 MPa (not shown in Fig. 4).
Figure 1. Schematic illustration of HPT device showing the compression and torsional stage. Specimens for light microscopy and microhardness measurements were first mechanically grinded on watered abrasive papers, then polished with polishing suspension of grade 3 and 1 um. Using this procedure, flat specimens with minimum surface scratches were obtained. Specimens for transmission electron microscopy were prepared by mechanical polishing followed by ion milling using PEPS ion mill at 4 kV. TEM observations were performed using Jeol 2000 FX electron microscope operated at 200 kV. Vickers microhardness (100g load) was measured on the semiautomatic Wolpert tester allowing automatic indentation. The regular square network of indents with the step of 0.5mm was performed in one quarter of each specimen after HPT. In order to find the exact center of the specimen two additional lines of indents on the other half of the specimen were also done. In Fig. 2 the optical micrograph showing the layout of individual indents on the specimen is shown. Using this procedure the center of each specimen was found with the accuracy of +/- 0.25mm.
By detailed inspection of Fig. 4 it is seen that the initially homogeneous distribution of a compressed specimen changes. HPT staining introduces an inhomogeneity in the material which is manifested by a clear minimum in Hv values in the center of the specimen after one turn (N=l). In this specimen the microhardness increases with increasing distance from the center reaching its maxium of 110 MPa which is approximately 30 MPa higher than in the center (Hv = 80 MPa). With increasing number of turns this difference gradually decreases by extension of the zone of maximum hardness from the periphery towards the center of the specimen. In the specimen after 5 turns (N=5) almost homogeneous distribution of microhardness was observed if the scatter of measured data is considered. In the specimen after 15 turns no differences in microhardness throughout the specimen were found. Note also that the maximum hardness does not change with increasing strain (number of turns). Within the experimental scatter the maximum value of approximately 112 MPa was observed in all specimens.
Figure 2. The square network of indents on the HPT specimen.
590
preliminary investigation may be summarized as follows: in the specimen after 1 turn significantly smaller grains were observed in the periphery of the disk than in the center. On the other hand, almost homogeneous distribution of grain sizes was observed in the specimen after 15 turns.
Microhardness results correspond well with our preliminary microstructure examinations by TEM. TEM proved significant grain refinement in specimens after HPT. One example of the microstructure is shown in Fig. 5. It displays the microstructure in the middle part of the specimen after one turn of HTP (s « 4). It is seen that HPT resulted in strong grain refinement. New grains of the average size of approximately 800 nm are clearly seen in the micrograph. Most grains contain many dislocations. However, several recrystallized almost dislocation free grains are also seen in Fig. 5. The individual grains have high misorientation as confirmed by the contrast or by detail electron diffraction analysis.
Changes of average Hv value in different parts of the disks are shown in Fig. 6 confirming the different ways towards the saturation in the center, in the middle section and at the periphery of the specimen.
£»!ri
aè»Àa r
• Jo
100
V
•o A
90
X
N N N N N
=1 =3 =5 = 15 =0
Periphery Middle Center
0
16 N - number of revolutinos
-10
Figure 6. The evolution of average microhardness values in different parts of the disks due to HPT straining.
0 r [mm]
Figure 4. Microhardness distribution across the diameters of AZ31 disks subjected to a pressure of 2.5 GPa and up to 15 whole number of revolutions.
In order to obtain more complex image of microhardness evolution throughout individual specimens subjected to different numbers of HPT rotations three-dimensional meshes of Hv data were constructed using the following procedure. The microhardness data were measured in one quarter of each disk as described in the previous section (approximately 400-500 indents were made in each specimen). First, these data were depicted, minimal microhardness found and matched with the center of the sample. Due to the step of measurement the center of the disk was found with the accuracy of +/- 0.25 mm. Data for the whole specimen were constructed from measured data symmetrically with respect to the center (data at the same distance from the center that were measured twice were both used and the average value calculated). However, such "completed data" are not suitable for 3D-depicting. The following smoothing procedure was applied to remove each wrongly indented or evaluated datum. Each value was recomputed using the original value and values of all close neighbors (4 edge neighbors and 4 corner neighbors). The distance of corner neighbors is V2 times higher than for edge neighbors, so the weight of corner neighbors is divided by V2 in the smoothing algorithm. Smoothing can be simply demonstrated by the following equation: FV=C.OV+
Figure 5. Typical microstructure of the specimen after 1 HPT rotation (middle section, e = 4).
°=y4 z
4 + 2v2 -"'
=
(
CN. EN. + yJ-
41
(1)
where FV is the final value, OV is the original value, EN is the edge neighbor, CN is the corner neighbor; fraction denominator represents partial normalization and finally c and b are smoothing parameters. However b and c are not independent (because of total normalization, i.e. b + c =1). Thus only e.g. c can be
Most grain boundaries are very close to equilibrium. It is manifested by a typical thickness band structure. The detailed analysis of TEM observations in individual parts of specimens is still in progress. However, general conclusions obtained from the
591
arbitrarily chosen. The choice c = 1 means that no smoothing occurs, whereas minimal meaningful value of c is 0.12. Otherwise the weight of each edge neighbor is higher than the weight of original values. In our case the value of c = 0.3 was used. It corresponds to the weight of 0.3, 0.1 and 0.07 for the original value, the edge neighbor and the corner neighbor, respectively.
Detail inspection of 3D meshes in Fig. 7 reveals the undulating character of microhardness variations across HPT disks indicating the manner in which the deformation develops in AZ31 alloy. This problem was recently addressed by strain gradient plasticity modeling introduced by Estrin et al. [31]. The authors consider the material in the form of a composite material comprising cell walls and cell interiors. The dynamic recovery in the former occurs by climbing of dislocations whereas in the latter it is the cross slip of dislocations which controls the rate of dynamic recovery. The cross slip in AZ31 is very difficult due to high separation width between partial dislocations. For a low stacking fault energy material the model predicts the behavior observed in Fig. 4 with clear minimum in the center of the disk in early stages of straining tending to saturate with continuing deformation.
Equation (1) holds only for interior points (a point that has all eight neighbors). Points that have incomplete number of neighbors are treated in the manner that nonexistent neighbors are ignored and weights of others are appropriately adjusted. Similarly, this procedure allows healing the missing data from the interior of the sample. The procedure extrapolates any missing value from values of the neighbors (if at least 5 of 8 possible neighbors exist). This procedure allows depicting 3D plots in a readable way even for partly damaged and/or missing data. Unlike Fig. 4 which is based on taking measurements of the variation in the microhardness values following diameters across disks after HPT processing, Fig. 7 shows integrated data across the total surface of individual disks. In Fig. 7 three-dimensional meshes of microhardness for specimens subjected to different number of turns are displayed. These plots were obtained by symmetrical completion of measured data and by smoothing procedure described in the previous paragraph. The variations of microhardness with position within the disk are clearly displayed at these meshes. The pronounced drop of microhardness in the center of the specimen after one turn is clearly seen in Fig. 7a. In the specimen after 3 turns (Fig. 7b) the central drop is still visible even if its depth is much lower in comparison with the specimen after 1 turn. The tendency to saturation with increasing number of turns is demonstrated at the mesh for the specimen after 15 turns (see Fig. 7c) whose surface is almost flat with slight undulations only, confirming saturated Hv values within the whole specimen. Discussion Microhardness Behavior and the Development of Homogeneity in AZ31 The variations in the values of microhardness across the diameters of AZ31 disks processed by HPT are in full agreement with the well-known dichotomic behavior reported by other authors in materials with face centered cubic structure and low stacking fault energy, e.g austenitic steel [26], Cu [27] and Ni [28]: i) lower hardness values were reported in the centers of disks and higher values in peripheral regions and ii) when torsional straining is continued to a sufficiently high total strain these variations tend to saturate and almost homogeneous microhardness distribution throughout the diameter of the disk is observed. Inspecting Fig. 6 shows that the saturation is achieved after 5 HPT rotations. As reported by Somekawa et al. [29] AZ31 alloy has a stacking fault energy of 27.8 mJm"2. Our measurements indicate that it is the stacking fault energy rather than the lattice structure which controls the microstructure evolution of the material which is macroscopically manifested by microhardness behavior. Contrary to these results, the centers of disks for pure aluminum exhibit higher values of Hv than the surroundings areas in early stages of torsional straining [30]. The authors explain this inverse behavior by exceptionally high rate of dynamic recovery in this material. Relatively low rate of dynamic recovery may be therefore expected in AZ31 alloy.
Figure 7. Three- dimensional meshes of microhardness as a function of the number of turns: a) N= 1 turn, b) N=3 turns, c) N=15 turns.
592
Whereas by additional 14 rotations the imposed strain in the center tripled with respect to the strain after one rotation.
Strain imposed by HPT Due to the variability in strain across the radius of torsionally strained specimen there were many attempts how to express the strain imposed by HPT in terms of the radius of the disk and the number of HPT rotations. Exhaustive discussion of the definition of strain imposed during HPT processing allowing calculation of different strain types (the true accumulated strain, shear strain, equivalent von Mises strain, the equivalent strain and the true strain) using different models can be found in the review by Zhilyaev [32].
- Periphery
The most precise quantity, taking directly into account the change of the specimen thickness during HPT processing, is the true strain £ defined as: £ = ln 1 +
2ïïN.r h
0
(2)
+ ln| *>
h [mm] 1 0.8
3
0.76
5
0.75
15
0.74
8
10
12
14
16
Conclusions
Table I. The dependence of thefinalthickness of HPT specimens on the number of turns.
1
6
Figure 8 Dependence of the true strain on the number of HPT rotations in various parts of the specimen.
The initial thickness h0 of samples before HPT was 1 mm. The final thickness h of samples depends slightly on the number of HPT turns. Table I summarizes the values of final thickness of specimens after different number of HPT turns obtained as the average of thickness measurements at four specimens in the same condition. As the volume of the sample does not change significantly, the final thickness decrease was caused by outward flow of material due to a quasi-constrained setup of HPT anvils.
0
4
No. of revolutions
where N is number of revolutions, r is radius of the specimen, h0 and h is the original and final thickness of the specimen, respectively. Thefirstterm of the equation (2) corresponds to the torsion of the specimen, whereas the second term refers to the compression strain due to thickness reduction.
N
2
The influence of HPT straining on microhardness evolution and microstructure in cast magnesium alloy AZ31 was investigated. The following conclusions may be drawn from this work. • HPT straining results in strong grain refinement. • HPT straining introduces a radial inhomogeneity in the material with a pronounced minimum in the center and the maximum near specimen periphery. • The inhomogeneity in Hv is continuously smeared out with increasing number of turns by continuous extension of the zone of maximum hardness towards the specimen center. • The integrated Hv data across the whole surface of HPT specimens revealed undulating character of microhardness variations across HPT disks. • The rate of dynamic recovery in AZ31 is very low due to its the low stacking fault energy. • Preliminary observations of microstructure evolution correspond well with microhardness behavior. Acknowledgements
The theoretical dependence of strain on the number of revolutions calculated according to eq. (2) is shown in Fig. 8. Radii under consideration are 10 mm (the periphery of the sample), 5 mm (middle section) and 0.25 mm (the center - it is the approximate value of the distance from sample center at which the closest microhardness value could have been measured). It is of particular interest that the strain saturates very soon (after 3 revolutions), especially for peripheral and middle regions. To illustrate this: another two revolutions (3 together) do not cause the strain of peripheral and middle section to rise for more than one third of the strain imposed after the first revolution. On the other hand, by second and third revolution the strain imposed in the center doubles. Similar relations hold between 3 and 15 revolutions. By straining the sample for additional 14 rotations (15 altogether) the imposed strain on the periphery and in the middle doubled only as compared to the strain after one rotation.
593
The authors acknowledge funding through the research program MSM 0021620834 of Ministry of Education of the CR. Partial support by GACR under grant 106/09/0482 is also gratefully acknowledged. J. Vrâtnâ acknowledges financial support by GAUK under the Grant 9594/2009 and by Grant SVV-261303. References 1. S.R. Agnew, J.A. Horton, T.M. Lillo and D.W. Brown, "Enhanced Ductility in Strongly Textured Magnesium Produced by Equal Channel Angular (ECA) Processing", Scripta Materialia, 50 (2004), 377-381. 2. Y. Estrin, S.B.Yi, H.G. Brokmeier, Z. Züberova, S.C. Yoon, H.S. Kim, R.J. Hellmig, "Microstructure, texture and mechanical properties of the magnesium alloy AZ31 processed by ECAP", International Journal of Materials Research, 99 (2008), 50-55. 3. R.B. Figueiredo, T.G. Langdon, "The characteristics of superplastic flow in a magnesium alloy processed by ECAP", International Journal of Materials Research, 100 (2009), 843846.
4. R.Z. Valiev, T.G. Langdon, "Principles of equal-channel angular pressing as a processing tool for grain refinement", Progress in Materials Science, 51 (2006), 881-981. 5. Y. Saito, N. Tsuji, H. Utsunomiya, T. Sakai, R.G. Hong, "Ultra-fine grained bulk aluminum produced by accumulative roll-bonding (ARB) process", Scripta Materialia, 39 (1998), 1221-1227. 6. N.A. Smirnova, V.l. Levit, V.l. Pilyugin, R.I. Kuznetsov, L.S. Davydova, V.A. Sazonova, "Evolution of the structure of f.c.c. single crystal subjected to strong plastic deformation", Fizika Metallov i Metallovedeniya, 61 (1986), 1170-1177. 7. V.M. Segal, "Materials processing by a simple shear", Mater. Sei. Eng., A 197 (1995), 157-164. 8. R.Z.Valiev, D.A Salimonenko, N.K. Tsenev, P.B. Berbon, T.G. Langdon, "Observations of high strain rate superplasticity in commercial aluminium alloys with ultrafine grain sizes", Scripta Materialia, 37 (1997), 1945-1950. 9. Z. Horita, M. Furukawa, M. Nemoto, A.J. Barnes, T.G. Langdon, "Superplastic forming at high strain rates after severe plastic deformation", Ada Materialia, 48 (2000), 3633-3640. 10. A. Yamashita, Z. Horita, T.G. Langdon, "Improving the mechanical properties of magnesium and magnesium alloy through severe plastic deformation", Materials Science and Engineering, A 300 (2001), 142-147. 11. Z. Horita, K. Matsubara, T.G. Langdon, "A two-step processing route for achieving a superplastic forming capability in dilute magnesium alloys", Scripta Materialia, 47 (2002), 255-260. 12. K. Matsubara, Y. Miyahara, Z. Horita, T.G. Langdon, "Developing superplasticity in a magnesium alloy through a combination of extrusion and ECAP", Ada Materialia, 51 (2003), 3073-3084. 13. Z. Horita, T.G. Langdon, "Microstructures and microhardness of an aluminum alloy and pure copper after processing by highpressure torsion", Materials Science and Engineering, A 410-411 (2005), 422-425. 14. Z. Horita, D.J. Smith, M. Nemoto, R.Z. Valiev, T.G. Langdon, "Observations of grain boundary structure in submicrometergrained Cu and Ni using high-resolution electron microscopy", Journal of Materials Research, 13 (1998), 446-450. 15. Y.H. Zhao, Y.T. Zhu, X.Z. Liao, Z. Horita, T.G. Langdon, "Influence of stacking fault energy on the minimum grain size achieved in severe plastic deformation", Materials Science and Engineering, A 463 (2007), 22-26. 16. A.P. Zhilyaev, G.V. Nurislamova, B.K. Kim, M.D. Barö, J.A. Szpunar, T.G. Langdon, "Experimental parameters influencing grain refinement and microstructural evolution during highpressure torsion", Ada Materialia, 51 (2003), 753-765. 17. Z. Horita, D.J. Smith, M. Furukawa, M. Nemoto, R.Z. Valiev, T.G. Langdon, "An investigation of grain boundaries in submicrometer-grain Al_Mg solid solution alloys using highresolution electron microscopy", Journal of Materials Research, 11(1996), 1880-1890. 18. G, Sakai, Z, Horita, T,G. Langdon, "Grain refinement and superplasticity in an aluminum alloy processed by high-pressure torsion", Materials Science and Engineering, A 393 (2005), 344353. 19. S. Dobatkin, E.N. Bastarache, G. Sakai, T. Fujita, Z. Horita, T.G. Langdon, "Grain refinement and superplastic flow in an aluminum alloy processed by high-pressure torsion", Materials Science and Engineering, A 408 (2005), 141-146. 20. Y.U. Ivanisenko, I. MacLaren, X. Sauvage, R.Z.Valiev, H.J. Fecht, "Shear-induced a —> y transformation in nanoscale Fe-C composite", Ada Materialia, 54(2006), 1659-1669.
21. S.S.M. Tavares, D. Gunderov, V. Stolyarov, J.M. Neto, "Phase transformation induced by severe plastic deformation in the AISI 304L stainless steel", Materials Science and Engineering, A 358 (2003), 32-36. 22. J. Petruzelka, L. Dluhoä, D. Hruäak, J. Sochova, "Nanostructure titanium - new material for dental implants (in Czech)", Ceskâ Stomatologickâ Rocenka, 106 (2006), 72-77. 23. L. Saldana, A. Mendez-Vilas, L. Jiang, M. Multigner, J.L. Gonzalez-Carrasco, M.T. Perez-Prado et al. "In vitro biocompatibility of an ultrafine grained zirconium", Biomaterials, 28 (2007), 4343-4354. 24. M. Kai, Z. Horita, T.G. Langdon, "Developing grain refinement and superplasticity in a magnesium alloy processed by high-pressure torsion", Materials Science and Engineering, A 488 (2008), 117-124. 25. Y. Harai, M. Kai, K. Kaneko, Z. Horita, T.G. Langdon, "Microstructural and Mechanical Characteristics of an AZ61 Magnesium Alloy Processed by High-Pressure Torsion", Materials Transactions, 49 (2008), 76-83. 26. A. Vorhauer, R. Pippan, "On the homogeneity of deformation by high pressure torsion," Scripta Materialia, 51 (2004), 921-925. 27. H. Jiang, Y.T. Zhu, D.P. Butt, I.V. Alexandrov, T.C. Lowe, "Microstructural evolution, microhardness and thermal stability of HPT - processed Cu", Materials Science and Engineering, A 290 (2000) 128-138. 28. Z. Yang, U. Welzel, "Microstructure-microhardness relation of nanostructured Ni produced by high-pressure torsion", Materials Letters, 59 (2005), 3406-3409. 29. H. Somekawa, K. Hirai, H. Watanabe, Y. Takigawa, K. Higashi, "Dislocation creep behavior in Mg-Al-Zn alloys", Materials Science and Engineering, A 407 (2005), 53-61. 30. C. Xu, Z. Horita, T.G. Langdon, "The evolution of homogeneity in processing by high-pressure torsion", Ada Materialia, 55 (2007), 203-212. 31. Y. Estrin, A. Molotnikov, C.H.J. Davies, R. Lapovok, "Strain gradient plasticity modelling of high-pressure torsion", Journal of the Mechanics and Physics of Solids, 56 (2008), 1186-1202. 32. A.P. Zhilyaev, T.G. Langdon, "Using high/pressure torsion for metal processing: Fundamentals and applications", Progress in Materials Science, 53 (2008), 893-979.
594
Magnesium Technology 2011 Edited by: Wim H. Sillekens, Sean R. Agnew, Neale R. Neelameggham, andSuveen N. Mathaudhu TMS (The Minerals, Metals & Materials Society), 2011
PLASTIC DEFORMATION OF MAGNESIUM ALLOY SUBJECTED TO COMPRESSION-FIRST CYCLIC LOADING Soo Yeol Lee , Michael A. Gharghouri , and John H. Root Department of Materials Engineering, The University of British Columbia, Vancouver, B.C. V6T 1Z4, Canada 2 Canadian Neutron Beam Centre, National Research Council Canada, Chalk River, ON KOJ 1J0, Canada Keywords: Magnesium, Plastic deformation, Twinning, Neutron diffraction Abstract
Experimental details
In-situ neutron diffraction has been employed to study the deformation mechanisms in a precipitation-hardened and extruded Mg-8.5wt.% Al alloy subjected to compression followed by reverse tension. The starting texture is such that the basal poles of most grains are oriented normal to the extrusion axis and a small portion of grains are oriented with the basal pole parallel to the extrusion axis. Diffraction peak intensities for several grain orientations monitored in-situ during deformation show that deformation twinning plays an important role in the elastic-plastic transition and subsequent plastic deformation behavior. Significant non-linear behavior is observed during unloading after compression and appears to be due to detwinning. This effect is much stronger after compressive loading than after tensile loading.
The binary alloy was supplied by the Péchiney Research Centre, Voreppe, France. The material was solution treated and aged. The volume fraction of ß-Mg17Al12 precipitates is -10%. The precipitates are plate shaped, with an average thickness of 165nm, a length that varies from 6 to 20p.m and a length to width ratio between 2 and 6. The grains are equiaxed with an average size of -60 urn. Neutron diffraction experiments were conducted on the L3 spectrometer of the Canadian Neutron Beam Centre, Chalk River Laboratories, Canada. Data were acquired for the (10-10), (10-11) and (0002) diffraction peaks with the scattering vector parallel to and normal to the applied load. The experiments were conducted in-situ, which allowed us to observe lattice strain evolution and bulk texture development as a function of applied load. The material microstructure and neutron diffraction experiments are described in detail elsewhere [11].
Introduction Magnesium and its alloy are currently the subject of many studies due to their potential use for lightweight structures in the automotive and aircraft industries, and for portable electronic devices [1,2]. The poor room-temperature formability of these alloys arises from the limited number of available slip systems [3,4]. The primary slip system in magnesium is slip with a l/3
Figure 1. (0002) pole figure showing initial texture of the test material measured by neutron diffraction. The pole figure is contoured in multiples of random distribution (m.r.d) with the thick solid black line corresponding to 1 m.r.d. The contour levels above and below 1 m.r.d are shown by solid and dotted lines, respectively, in 0.5 m.r.d steps.
In this work, lattice strains and diffraction peak intensities for several grain orientations were measured in an extruded Mg8.5wt.% Al alloys during deformation using neutron diffraction. The loading path consisted of compression followed by reverse tension.
Results and Discussion The (0002) basal pole figure for the as-received material is shown in Fig. 1. The center of the pole figure corresponds to the extrusion axis. The (0002) basal planes in most grains are parallel
595
or close to the extrusion axis, with a correspondingly small fraction of grains with the basal planes perpendicular to the extrusion axis. Fig. 2 shows the macroscopic stress-strain response of the alloy subjected to compression followed by reverse tension. The symbols in the graph correspond to points in the loading history at which diffraction data were acquired. In compression up to -100 MPa, the alloy shows mainly elastic deformation, with some limited plasticity. Beyond -100 MPa, the elastic-plastic transition is well underway, though plasticity is not fully developed. During unloading after compression and reloading in tension, the stress-strain response is clearly nonlinear, showing a typical s-shaped profile. The material undergoes general yielding in tension at —1-125 MPa. The unloading portion of the curve, after tensile loading, is again non-linear, but the effect is less significant than after compression.
1
1
■
1
'
1
a a.
I
150
i
, •■
100 50 0 -50
/
O)
c -100
111
-150 -0.8
•/
. •/ ; / <• J.L • TÎ * i /
/• /• • • • • •
■
i
-0.4
i
•/ •/
As the load increases from -100 MPa to -130 MPa, the (0002) intensity increases sharply due to {10-12}<10-11> extension twinning in grains with the c-axis normal to the loading direction, such as the {10-10} grains. It is thus likely that this extension twinning is responsible for the observed plastic deformation behavior shown in Fig. 2. Since many grains in the starting material have the c-axis approximately normal to the loading direction, the lattice reorientation due to extension twinning will reorient the material into the (0002) orientation. The effect on the intensity of the (0002) reflection is thus very strong, resulting in a 3-fold increase even at very low macroscopic strain.
__•
■
200 -Compression-first cyclic loading
!
if material is re-orientated. In the present case, no phase change occurs. Thus, a decrease in intensity of a given reflection means that the contributing grains have twinned - i.e. that some of the grain volume is no longer appropriately oriented for diffraction. Conversely, an increase in intensity means that a conjugate set of grains has undergone twinning, such that additional material contributes to the intensity. In the case of magnesium, {10-12} extension twinning results in an 86.6° reorientation of the crystal lattice, which would convert an {10-10} orientation into an (0002) orientation and vice versa.
^ - ^^^^ •
•
/
•/ / • •/ / /• •
•/ ■ i i
■
i i i
0.0
i
i
0.4
i
i
0.8
1.2
1.6
Engineering Strain (%)
Figure 2. Macroscopic stress-strain response of the extruded Mg8.5wt.%Al alloy. The symbols correspond to points in the loading history at which diffraction data were acquired. The integrated intensity data acquired during loading for all measured reflections are shown in Fig. 3. For these measurements, the scattering vector was parallel to the loading direction. The data are plotted chronologically as functions of the applied load in order to show clearly how the measured values change during the test. The (0002), {10-11} and {10-10} orientations have the basal poles (c-axes) oriented at 0°, 61.9° and 90°, respectively, relative to the extrusion direction (ED), which is also the loading direction. Thus, the {10-10} and (0002) grains are favorably oriented for extension twinning in compression and tension, respectively. Neither family of grains is oriented favorably for -slip on any plane. In contrast, the {10-11} grains are favorably oriented for basal slip. The {411} reflection is from the ß-Mg17Ali2 precipitates.
Figure 3. (a) Intensity variations as functions of applied load during the deformation shown in Figure 2; (b) Close-up of intensity variations for the (0002) reflection. Measured (hkit) plane normal is parallel to the applied loading direction, which is the same as the extrusion direction (ED).
During initial loading in compression, the (0002) intensity starts to increase beyond —100 MPa, while the {10-10} intensity decreases concurrently. This corresponds with the macroscopic yield point (—100 MPa). The intensity of a diffraction peak changes when the amount of material oriented for Bragg diffraction changes. Since the volume sampled by the neutron beam is constant, this can only occur if there is a phase change, or
When the sample is unloaded after compression, the intensity of the (0002) peak remains stable down to an applied stress of —50 MPa, and then decreases gradually. The intensity falls by about 40% during unloading to zero load, indicating that about 40% of the twinned volume has detwinned during unloading after compression. It is possible that the parent (0002) grains undergo compressive {10—11 }<10—12> twinning during unloading after
596
compression, but this is unlikely as this type of twinning is very uncommon, and is generally found to contribute little to lattice reorientation during deformation. This detwinning behavior is thought to contribute significantly to the non-linear behavior observed in Fig. 2.
25(2009)861-880. 10. O. Murânsky, M.R. Barnett, V. Luzin, S. Vogel, Mater. Sei. Eng. A 527 (2010) 1383-1394. 11. M.A. Gharghouri, G.C. Weatherly, J.D. Embury, J. Root, Phil. Mag. A 79 (1999) 1671-1695.
During reverse loading in tension, detwinning continues until the (0002) intensity at the start of the test is fully recovered at about + 100 MPa. As the applied load increases from 100 MPa to 205 MPa, the (0002) intensity continues to decrease, but at a lower rate. At this point, it appears that the (0002) minority grains undergo {10-12} extension twinning. During unloading after the tensile portion of the stress-strain curve, the (0002) peak intensity does not change, suggesting that the twinned material in the minority (0002) grains does not undergo significant detwinning.
12. D.W. Brown, S.R. Agnew, M.A.M. Bourke, T.M. Holden, S.C. Vogel, C.N. Tome, Mater. Sei. Eng. A 399 (2005) 1-12. 13. L. Wu, S.R. Agnew, D.W. Brown, G.M. Stoica, B. Clausen, A. Jain, D.E. Fielden, P.K. Liaw, Acta mater. 56 (2008) 3699-3707. 14. O. Murânsky, M.R. Barnett, D.G. Carr, S.C. Vogel, E.C. Oliver, Acta mater. 58 (2010) 1503-1517.
Summary Neutron diffraction has been used to study the plastic deformation behavior of a Mg-8.5wt.%A1 alloy subjected to compression followed by reverse tension. It was found that the onset of extension twinning corresponds well with the macroscopic elasticplastic transition. The non-linear behavior during unloading after compression was more significant than that after tension. Diffraction results show that about 40% of the twinned volume detwins during unloading after compression, but that no significant detwinning occurs during unloading after tension. Acknowledgements This work was supported by funding from the NSERC Magnesium Strategic Research Network. More information on the Network can be found at www.MagNET.ubc.ca. References 1.
M.M. Avedesian and H. Baker, Magnesium and Magnesium Alloys, ASM Specialty Handbook, ASM International, Materials Park, OH, 1999.
2.
C. Wang, P. Han, L. Zhang, C. Zhang, X. Yan, B. Xu, J. All. Comp. 482 (2009) 540-543.
3.
C.S. Roberts, Magnesium and its Alloys, John Wiley & Sons, Inc., New York, 1960.
4.
O. Murânsky, D.G. Carr, M.R. Barnett, E.C. Oliver, P. Sittner, Mater. Sei. Eng. A496 (2008) 14-24
5.
P.G. Partridge, Metall. Rev. 12 (1967) 169-194.
6.
S.R. Agnew, C.N. Tome, D.W. Brown, T.M. Holden, S.C. Vogel, Scripta mater. 48 (2003) 1003-1008.
7.
J. Jain, W.J. Poole, C.W. Sinclair, M.A. Gharghouri, Scripta mater. 62(2010)301-304.
8.
B. Clausen, C.N. Tome, D.W. Brown, S.R. Agnew, Acta mater. 56 (2008) 2456-2468.
9.
G. Proust, C.N. Tome, A. Jain, S.R. Agnew, Int. J. Plasticity
597
Magnesium Technology 2011 Edited by: Wim H. Sillekens, Sean R. Agnew, Neale R. Neelameggham, andSuveen N. Mathaudhu TMS (The Minerals, Metals & Materials Society), 2011
MICROSTRUCTURE EVOLUTION IN AZ61L DURING T T M P AND SUBSEQUENT ANNEALING TREATMENTS T.D. Berman 1 , W. Donlon 1 , R. Decker 2 , J. Huang 2 , T.M. Pollock 3 , J.W. Jones 1 Materials Science and Engineering, University of Michigan, 2300 Hayward St; Ann Arbor, MI, 48109, USA 2 nanoMAG, L L C , 620 Technology Drive; Ann Arbor, MI, 48108, USA 3 Materials Department, University of California, Santa Barbara; Santa Barbara, CA, 93106-5050, USA Keywords: AZ61, Microstructures, Texture, Mechanical Properties, Thixomolding, Thermomechanical Processing Table I: Composition of received AZ61L in wt %
Abstract Microstructure evolution is studied in Thixomolded® Thermomechanical Processed (TTMP) AZ61L sheet at various stages of processing. Transmission electron microscopy (TEM) is utilized to examine (1) grain refinement and recrystallization and (2) refinement and re-distribution of the /3-Mg 17 Al 12 phase in the as-Thixomolded, as-TTMP, and annealed conditions. Electron backscatter diffraction (EBSD) is used to study texture evolution through T T M P and annealing. The influence of microstructure produced by T T M P and annealing on the mechanical properties will be discussed.
Al 6.5
Zn 0.46
Mn 0.14
Si 0.01
Fe 0.003
Mg bal.
Dogbone tensile specimens with a gauge length of 31.75 mm and cross section of 7.94 mm, were machined from the sheets with the tensile axes parallel to the rolling direction. Room temperature tensile tests were performed with a displacement rate of 0.71 mm/min. An extensonometer was used to measure tensile elongation. Samples for microscopy were removed from the grip ends of the tensile bars in the plane of the sheet. TEM specimens were prepared by electropolishing with a 8% perchloric acid in methanol electrolyte or dimpled and ion milled using a Gatan Precision Ion Polishing System (PIPS). TEM was conducted with a JEOL2000FX and a Phillips CM12 AEM systems. EBSD specimens were prepared by polishing with 1 /xm diamond followed by ion polishing in a Gatan PIPS. EBSD examination was conducted on a JEOL 840A SEM system equipped with a HKL EBSD system. A step size of 0.5 /un was used.
Introduction The high cost and poor formability of magnesium sheet has limited its commercial application [1]. If, through materials design and processing, formability could be enhanced, several markets, especially automotive, would benefit. The emerging consensus is that deformation modes required for high ductility materials are more easily activated in fine, micro- or nanoscale grains [2]. Several different processes to achieve fine scale microstructure have been developed, producing high strength, high ductility materials [3-10]. Thixomolding Thermal Mechanical Processing (TTMP) builds upon the fine grains, isotropy, and low porosity of Thixomolded Mg alloys containing eutectic phases. Intense thermomechanical processing is applied to further refine grain size and eutectic phases. Furthermore, additional thermal treatments can be applied to optimize strength, ductility and formability.
Results
Experimental
The as-Thixomolded material reveals equiaxed grains, regions of high dislocation density, and both agglomerated (indicated by the arrow) and isolated ß particles (Fig. 1). In addition to agglomerated and isolated, equiaxed ß grains, regions of eutectic ß are also found in the as-Thixomolded condition. Fig. 2 highlights a region of eutectic ß, likely composed of a series of small ß grains along a boundary.
Thixomolded and T T M P samples of AZ61L were obtained from Thixomat, Inc., composition shown in Table I. A rolling pass yielded a sheet thickness of 1.5 mm for the T T M P samples, a 50% reduction in the thickness of the as-Thixomolded plates. Annealing temperatures of 250 °C and 300 °C were chosen to explore effects of annealing on microstructure and mechanical properties.
The TTMP-induced microstructure demonstrates significant grain refinement from the Thixomolded condition (Fig. 3). Highly elongated,deformed grain structures are seen (region A) in the T T M P condition. A collection of nanoscale grains, that may indicate local dynamic recrystallization occurred, are shown in region B. The elongated ß structures and clumps of ß particles
599
Figure 1: Microstructure of Thixomolded AZ61L. The arrow highlights a region of agglomerated ß particles.
Figure 3: Elongated grains (A) and nanoscale a-Mg and ß grains (B) are found in the as-TTMP condition.
observed in the as-Thixomolded material are dissipated. ß grains are heterogeneously distributed throughout the as-TMP microstructure.
ß particles were frequently observed in the interior of grains and along a-Mg grain boundaries. Nano-scale recrystalized a-Mg grains tend to coincide with the ß phase. Clusters of ß particles and recrystallized a-Mg grains suggests that the ß particles play a significant role in retarding grain growth.
Annealing treatments allow for further recrystallization. At 250 °C the material initially retains a fine grain microstructure. Grain coarsening becomes evident after 20 minutes at 250°C (top row Fig. 4). Annealing at 300 °C initiates more rapid grain growth, evident even at the shortest annealing time (bottom row Fig. 4). Heterogeneous distribution of ß grains is evident in Figs. 4(c-e). Nano-scale ß particles in the TMP + annealed material were equiaxed and often exhibited a rounded morphology as can be seen in Fig. 5 and Fig. 6.
Figure 2: TEM micrograph showing the morphology of the eutectic ß particles in the as-Thixomolded condition. Identification of phase was determined by EDAX measurements.
A subset of the samples are analyzed by EBSD. As-TTMP and T T M P samples annealed for short times exhibit many (~75%) non-indexable EBSD diffraction patterns due to high deformation. The TMP samples annealed for longer durations yield fewer (~10%) non-indexable patterns. Band contrast images, with re-constructed grain boundaries (>5% misorientation), are shown in Fig. 7. Finer grains with a bimodal size distribution is evident for the T T M P + 300 °C 20 min anneal sample, demonstrating the effect of recrystallization. The maximum grain diameters for the as-Thixomolded and T T M P + 300 °C 20 min anneal are 36 ^m and 14 ßm respectively. Pole figures were also obtained from the EBSD analysis. Nearly random texture patterns are observed in the as-Thixomolded and T T M P + 300 °C 20 min anneal sample (Fig. 8(a and d)). Strong (0001) texture patterns were observed for the as-TTMP sample(Fig. 8(b)) and the T T M P + 250 °C samples annealed for short times (Fig. 8(c)). In addition to the (0001) component, a weaker (1100) texture was observed to follow the same trend parallel to the rolling direction.
F i g u r e 4: Microstructure evolution resulting from post T T M P annealing treatments (a) T T M P + 250 °C 3 min anneal, (b) T T M P + 250 °C 10 min anneal, (c) T T M P + 250 °C 20 min anneal, (d) T T M P + 300 °C 3 min anneal, (e) T T M P + 300 °C 10 min anneal, (f)TTMP + 300 °C 20 min anneal.
F i g u r e 5: T E M micrograph of a sample annealed at 250 °C for 20 min with small a-Mg grains and ß grains. Identification of the ß phase was determined by EDAX measurements.
F i g u r e 6: A larger ß particle (dark) in a cluster of smaller, presumably ß particles in the sample annealed at 300 °C for 20 min. SAED pattern from the [Ï13] zone axis of the ß crystal.
601
Table II: Mechanical properties of the AZ61L samples examined in this study Sample
YS (MPa)
U T S (MPa)
as-Thixomolded as-TTMP T T M P + 250 °C T T M P + 250 °C T T M P + 250 °C T T M P + 300 °C T T M P + 300 °C T T M P + 300 °C
127.3 ± 315.8 ± 331.5 ± 313.0 ± 321.7 ± 236.7 ± 223.6 ± 216.5 ±
191.5 ± 370.7 ± 372.3 ± 365.3 ± 361.3 ± 315.3 ± 308.5 ± 307.6 ±
3 min Anneal 10 min Anneal 20 min Anneal 3 min Anneal 10 min Anneal 20 min Anneal
14.7 6.8 6.4 17.2 3.1 4.1 1.2 5.7
16.1 4.3 5.4 8.0 2.3 1.3 4.8 1.4
El (%) 5.1 ± 6.7 ± 5.5 ± 5.6 ± 7.2 ± 15.5 ± 16.6 ± 22.5 ±
0.4 3.9 2.4 3.4 1.4 2.1 3.0 1.5
Figure 8: (0001) Pole figures of (a) as-Thixomolded, (b) asTTMP, (c) TTMP + 250 °C 3 min anneal, and (d) TTMP + 300 °C 20 min anneal. Mapping was performed in the plane of the sheet with the rolling direction pointing up.
Discussion It has previously been demonstrated that the effect of TMP on Thixomolded AZ61L is a marked increase in the yield strength and ultimate tensile strength of the material, without reduction of tensile ductility [3, 11, 12]. A summary of the mechanical characteristics of the material can be found in Table II. The TMP approximately doubled yield and tensile strength over that of the as-Thixomolded sample. Annealing at 250 °C had little affect on the tensile properties. As grains coarsen during the 300 °C annealing treatments, yield and tensile strength decreases . For the times and temperatures selected, annealing temperature had a m o r e significant affect on tensile properties than annealing duration. An increase in ductility of samples annealed at 300 °C may result from a loss of deleterious texture as recrystallization occurs.
Figure 7: EBSD maps of (a) as-Thixomolded and (b) TTMP + 300 C 20 min anneal
602
Low R value measurements taken in similarly processed AM60, indicate a decreased anisotropy from the asT T M P condition, corresponding with the demonstrated loss of texture after the 300 °C 20 min anneal [11]. Texture evolution and R value measurements in T T M P AZ61L with anneal and/or aging treatments will be an area of future exploration. To our knowledge, non-textured magnesium sheet has not been seen in conventional rolled material.
4. J. Swiostek et al., "Hydrostatic Extrusion of Commercial Magnesium Alloys at 100 °C and its Influence on Grain Refinement and Mechanical Properties," Mater Sei Eng A, 424 (1-2) (2006), 223-229. 5. K. Matsubara et al., "Developing Superplasticity in a Magnesium Alloy Through a Combination of Extrusion and ECAP," Ada Mater, 51 (11) (2003), 3073-3084. 6. D. H. St. John et al., "Grain Refinement of Magnesium Alloys II . Grain Refinement of Magnesium Technical Status," Metall Mater Trans A, 36 (July) (2005), 1669-1679.
Future investigations will further explore methods to (1) minimize texture after TTMP, (2) control and refine the distribution of ß particles, and (3) optimize the heattreatment to better control grain size, optimizing the mechanical properties.
7. H. S. Di et al., "New Processing Technology of Twin Roll Strip Casting of AZ31B Magnesium Strip," Mater Sei Forum, 488-489 (2005), 615-618.
Conclusions
8. S. M. Zhang, Z. Fan, and Z. Zhen, "Direct Chill Rheocasting (DCRC) of AZ31 Mg Alloy," Mater Sei Technol, 22 (12) (2006), 1489-1498.
TMP of Thixomolded AZ61L results in very fine grain sheet with a twofold increase in strength with no loss of ductility. Annealing at 250 °C had little affect on the mechanical properties from the as-TTMP condition and evidence of grain growth is not seen until 20 minutes. Grain coarsening and a drop in strength is seen after annealing the T T M P material at 300 °C. Annealing at 300 °C for 20 min was shown to lead to a nearly non-textured sheet with ductility several times greater than in the As-TTMP material. These results show the promise of optimizing the T T M P and post treatments to develop fine grain, non-textured magnesium sheet with high strength and formability.
9. W.-J. Kim, G. E. Lee, and J. B. Lee, "Achieving Low Temperature Superplasticity from CaContaining Magnesium Alloy Sheets," Adv Eng Mater, 11 (7) (2009), 525-529. 10. G. Cao et al., "Study on Tensile Properties and Microstructure of Cast AZ91D/A1N Nanocomposites," Mater Sei Eng A, 494 (1-2) (2008), 127-131. 11. R. Decker et al., "Thixomolded and Thermomechanically Processed Fine-Grained Magnesium Alloys," Mater Sei Forum, 654-656 (2010), 574-579.
Acknowledgement
12. J. Huang et al., "On Mechanical Properties and Microstructure of T T M P Wrought Mg Alloys," Magnesium Technology 2010, eds. S. Agnew et al. (Minerals, Metals, and Materials Society, Warrendale, PA, 2010), 489-493.
This material is based upon work supported by the National Science Foundation under Grant No. 0847198. References 1. S. Mathaudhu and E. Nyberg, "Magnesium Alloys in U.S. Military Applications: Past, Current and Future Solutions," Magnesium Technology 2010, eds. S. Agnew et al. (Minerals, Metals, and Materials Society, Warrendale, PA, 2010), 27-33. 2. J. Koike, "Enhanced Deformation Mechanisms by Anisotropie Plasticity in Polycrystalline Mg Alloys at Room Temperature," Metall Mater Trans A, 36 (7) (2005), 1689-1696. 3. R. Decker et al., "nanoMAG® High Strength/Density Mg Alloy Sheet," Magnesium Technology 2009, eds. E. Nyberg et al. (Minerals, Metals, and Materials Society, Warrendale, PA, 2009), 489-493.
603
Magnesium Technology 2011 Edited by: Wim H. Sillekens, Sean R. Agnew, Neale R. Neelameggham, andSuveen N. Mathaudhu TMS (The Minerals, Metals & Materials Society), 2011
MODELING THE CORROSIVE EFFECTS OF VARIOUS MAGNESIUM ALLOYS EXPOSED TO TWO SALTWATER ENVIRONMENTS H.J. Martin, C. Walton, J. Danzy, A. Hicks, M.F. Horstemeyer, P.T. Wang Center for Advanced Vehicular Systems (CAVS), Mississippi State University, Box 5405, Mississippi State, MS 39762, USA Keywords: Magnesium Alloy, Pitting Corrosion, Intergranular Corrosion, Modeling of Corrosion The presence of alloying elements is not the only consideration when dealing with corrosion. The formation method, whether it is extrusion or casting, plays a significant role in the corrosion properties of magnesium alloys. Casting results in the formation of an as-cast "skin", with very small grains, which increases the corrosion resistance of magnesium, up to ten-fold higher than the bulk material [9, 14]. Since extrusion removes this as-cast skin, and does not allow the formation of another skin, extrusion can negatively affect the corrosion resistance of magnesium.
Abstract The use of magnesium within the automotive industry is limited by its corrosion rate in the presence of saltwater. By adding various elements, the magnesium microstructure and corrosion rate can be altered. In the Center for Advanced Vehicular Systems at Mississippi State University, a model is being developed to elucidate the total corrosion of magnesium alloys and is comprised of general corrosion and pitting corrosion, respectively, as shown below:
t = ^ c H he
While the addition of elements to magnesium can increase the corrosion resistance of magnesium, currently, there are no available models to determine if the addition of certain elements will improve corrosion resistance. Models designed for stainless steel and aluminum attempt to quantify the causes of pitting, such as diffusion, energy, pH, and corrosion potential, but involve the use of a current to initiate the corrosion [15-22]. In addition, the only available model for magnesium requires a current to produce a polarization curve to predict galvanic corrosion, which is not applicable to this research [23]. The goal of this research, then, is to study various magnesium alloys in as-cast or extruded form in order to develop a model that accurately describes pit nucleation, pit growth, and pit coalescence.
(D
where pitting corrosion is based on the pit number density, pit surface area, and a nearest neighbor distance function, respectively, as shown below: hc=1
PVPC
(2)
The exposure environment resulted in differences in the amount of pit nucleation, in the size of the pits formed, and in the rate of coalescence. Time also affected the surface characteristics, as general corrosion began degrade the number and size of the pits. The research presented here will cover the model development, calibration, and validation.
Materials and Methods
Introduction Testing
Magnesium has a high corrosion rate as compared to aluminum or steel, relegating its role in the automotive and aerospace industries in places that are unexposed to the environment [1-3]. In an effort to improve the corrosion resistance of magnesium, various elements are added, including aluminum, zinc, manganese, and rare earth elements [2, 4-8].
Twelve AZ61 coupons (2.54 cm x 2.54 cm x varying thicknesses) were cut from an extruded crash rail provided by Ford using a CNC Mill (Haas, Oxnard, CA). Twelve AZ31 coupons ( cm x cm x cm) were cut from extruded sheets using a vertical band saw (MSC Industrial Supply Company, Columbus, MS). Twelve AM60 coupons (2.54 cm x 2.54 cm x varying thicknesses) were cut from as-cast control arms using a vertical band saw. Twelve AE44 coupons (2.54 cm x 2.54 cm x varying thicknesses) were cut from an as-cast engine cradle provided by Meridian Technologies using a vertical band saw. The coupon surfaces were left untreated to test the corrosion effects on the extruded AZ61 and AZ31 magnesium alloys and on the as-cast AM60 and AE44 magnesium alloys.
The addition of aluminum, up to 10%, has been shown to affect the corrosion resistance of magnesium [2]. Aluminum, which is present in the ß-phase precipitate, appearing as Mg17Al12, leads to improved corrosion resistance when the ß-phase is continuous and finely divided [2, 9-11]. However, the same ß-phase precipitate can lead to the creation of micro-galvanic cells, thereby decreasing the corrosion resistance, when the ß-phase is small and unconnected [2,9-11].
Two different test environments were used in this study: salt spray testing and immersion. For salt spray testing, a Q-Fog CCT (QPanel Lab Products, Cleveland, OH) was used to cycle through three stages set at equal times, including a 3.5 wt.% NaCl spray at 35°C, 100% humidity using distilled water at 35°C, and a drying purge at 35°C. For immersion testing, an aquarium with an aeration unit was filled with 3.5 wt.% NaCl at room temperature. For both tests, the six coupons per test environment were hung at 20° to the horizontal, as recommended by ASTM B-l 17 [24]. The coupons were exposed to the test environment for 1 h, removed, rinsed with distilled water to remove excess salt, and dried. Following the profilometer analysis, the coupons were then placed
In addition to the addition of aluminum, the presence of rare earth elements can affect the corrosion properties and mechanical properties of magnesium [4-8]. The formation of meta-stable rare earth element - containing phases along the grain boundaries improves the creep properties of magnesium and can also improve corrosion resistance due to trace amounts of the rare earth elements in the passive films formed during atmospheric exposure [6-8, 12]. The addition of the rare earth elements also shifts the location of pitting corrosion, from along the magnesium grain eutectic boundary to the interior of the magnesium grain, resulting in unaffected rare earth intermetallic regions [7, 13].
605
compared. As with pit number density, the most surfaces followed a second-order polynomial, with the highest pit area occurring on the AE44 immersion surface. Also notice that the AZ31 surface was divided by 10 so as not to compress the data on the the y-axis, making the three other surfaces indistinguishable. In addition, there was no decrease in pit area on the AZ61 salt spray surface or either AZ31 surfaces, which increased until the end of the experiment. The smallest pit area occurred on the AM60 as-cast surfaces.
back into the test environment for an additional 3 h, an additional 8 h, an additional 24 h, and another 24 h. These times allowed for a longitudinal study to follow pit growth and surface changes over time, where to = 0, ti = 1 h, t2 = 4 h, t3 = 12 h, U = 36 h, and t5 = 60 h. Between analyses and environmental exposures, the coupons were stored in a desiccator to ensure that no further surface reactions occurred. Analysis Following each time exposure, the coupons were analyzed using optical microscopy and laser profilometry. The coupons were weighed prior to testing and following each exposure on two different scales and an average was taken. Four thickness measurements were taken on each sample prior to and following the test. Because the coupons were cut from an engine cradle, the thicknesses of the coupons varied from side to side, meaning an average was taken per coupon based on the four measurements. Measurements for all figures were averaged from the data with error bars based on one standard deviation. Optical microscopy with an inverted light was used to take multiple images of the resulting surface at 5x magnification and lOx magnification (Axiovert 200M Mat, Carl Zeiss Imaging Solutions, Thornwood, NY). The 5x magnification images were combined and then analyzed using the ImageAnalyzer (v. 2.1-2) provided by Mississippi State University to determine the number of pits, the pit surface area, the nearest neighbor radius, and the intergranular corrosion area fraction necessary for the development of a corrosion model not detailed in this paper but previously outlined by Horstemeyer et al. [25]. The 10x magnification was used to pictorially show the changes over the six cycles. Laser profilometry was used to scan a 1 mm by 1 mm area on two coupons per environment following each test cycle (Talysurf CLI 2000, Taylor Hobson Precision Ltd, Leicester, England). The resulting 2-D and 3-D images were used to document the changes in the pit characteristics due to the different test environments over the six cycles (Talymap Universal, v. 3.18, Taylor Hobson Precision Ltd, Leicester, England). Data was collected based on fourteen pits within each 1 mm by 1 mm area, for a total of twenty-eight data points per environment per cycle.
Figure 1 : Average weight loss of various magnesium alloys based on test environment over 60 h. Notice that all surfaces followed logarithmic trends.
Results Figures 1 and 2 show the average weight and thickness loss, respectively, over the five exposure times for the immersion and salt spray surfaces on the various magnesium alloys being compared. As one can see, all surfaces follow similar logarithmic trends for weight loss (Figure 1) and thickness loss (Figure 2).
Figure 2: Average thickness loss of various magnesium alloys based on test environment over 60 h. Notice that all surfaces followed logarithmic trends.
Figure 3 shows the pit number density over the five exposure times for the immersion and salt spray surfaces on the various magnesium alloys being compared. As one can see, all surfaces followed second-order polynomial trends. The AZ61 surfaces showed the highest amount of pit formation as compared to the other surfaces, while the as-cast AM60 surfaces showed the lowest amount of pit formation. In addition, all immersion surfaces had higher pit number densities as compared to the respective salt spray surface.
Figure 5 shows the changes in the nearest neighbor distance, which is the distance between two pits, for the immersion and salt spray surfaces on the various magnesium alloys being compared. As with the pit number density and the pit area, the surfaces followed second-order polynomial trends, although in reverse of the pit number density and pit area. The extruded AZ61 surfaces showed the smallest nearest neighbor distance while the as-cast AM60 surfaces showed the largest nearest neighbor distance.
Figure 4 shows the changes in the pit area, which is the 2-D area covered by the pits as seen by micrographs for the immersion and salt spray surfaces on the various magnesium alloys being
Figure 6 shows the intergranular corrosion area fraction (ICAF), which is the fraction of the surface that shows the corrosion that
606
occurs in the ß-phase precipitate phase of the alloy, for the immersion and salt spray surfaces on the various magnesium alloys being compared. All surfaces follow logarithmic trends, with the highest ICAF occurring on the as-cast AE44 surfaces.
Figure 5: Nearest neighbor distance of various magnesium alloys based on test environment over 60 h. Notice that all surfaces followed second-order polynomial trends. Also notice that the ascast AM60 surfaces had the largest nearest neighbor distance while the extruded AZ61 surfaces had the smallest nearest neighbor distance.
Figure 3: Pit number density of various magnesium alloys based on test environment over 60 h. Notice that all surfaces followed second-order polynomial trends. Also notice that all immersion surfaces had higher pit number densities as compared to the respective salt spray surface.
Figure 6: Intergranular corrosion area fraction of various magnesium alloys based on test environment over 60 h. Notice that all surfaces followed logarithmic trends. Also notice that the as-cast AE44 surfaces had the largest intergranular corrosion area fraction.
Figure 4: In-plane pit area of various magnesium alloys based on test environment over 60 h. Notice that all surfaces followed second-order polynomial trends. Also notice that the extruded AZ31 surfaces had the largest pit area, which was divided by 10 to ensure all data could be seen. Notice also that the as-cast AM60 surfaces had the smallest pit area. In addition, there was no decrease in pit area on the extruded AZ61 salt spray surface.
General corrosion (
Discussion Total corrosion includes general corrosion, which occurs when water reacts with a magnesium surface to create a Mg(OH)2 film and H2 gas, and pitting corrosion, which occurs when chloride ions from salt water initiate and maintain pit formation on the magnesium surface [3]. At Mississippi State University, a model is currently being developed that incorporates general corrosion and pitting corrosion, as shown below: ■■Uc^ h
he
=1
PVPC
(2)
where T| is the pit number density, v is the pit area, and c is a function of the nearest neighbor distance and the intergranular corrosion area fraction (ICAF).
(1)
607
When looking at general corrosion, more weight loss is seen on the immersion surfaces as compared to the salt spray surfaces, except with respect to the as-cast AE44 surfaces (Figure 1). Because the samples in the immersion environment are continuously surrounded by salt containing water, more water can react with the surface, leading to more weight loss as magnesium is removed from the surface. When looking at the salt spray surfaces, which are exposed first to salt containing water, then to 100% humidity, and then to a dry phase, weight loss is not as significant, since there is no continuous exposure to water, meaning there is no continuous general corrosion ongoing. However, more thickness loss is seen on the salt spray surfaces as compared to the immersion surfaces (Figure 2). This is not expected, as one would expect that thickness loss would follow weight loss. The difference in thickness and weight loss, though, can be attributed to the way measurements were taken. Weight loss used a scale, meaning the entire coupon was measured, while thickness loss was taken using calipers, meaning only the edges of the coupons were measured. This difference can account for the difference in thickness loss, as cleaning the samples after each corrosion period removed significant pit debris and salt on the salt spray samples. While both samples hung at 20°, the samples exposed to the cyclical salt spray experienced a collection of pit debris and salt along the edges of the samples, due to the drying phase. This debris led to higher pitting corrosion along the edges, which were measured with the calipers. Because there was no collection of salt or pit debris along the edges of the immersion samples, extra pits could not form meaning the thickness was unaffected by the debris.
only pit when the chloride ions are present, resulting in lower numbers of pits forming. When comparing the form of magnesium, one can see that the extruded magnesium experienced much higher amounts of pit nucleation than the as-cast magnesium. This is due to the presence of an as-cast skin increasing the corrosion resistance on the as-cast alloys [9, 14]. This skin was removed during the extrusion process, meaning that the extruded magnesium can more easily corrode. Finally, the type of magnesium places a role in the formation of pitting. The higher the amount of aluminum, up to 10%, the higher the corrosion resistance [2]. When comparing the two as-cast materials, the lower pit nucleation corresponded with the higher percentage of aluminum, meaning that aluminum played a higher role in corrosion resistance than either zinc or rare earth elements. The differences seen in pit area (Figure 4) can also be attributed to the environment to which the magnesium is exposed, the form of magnesium, and the type of magnesium alloy. When looking at the environments, larger pits were seen on the AZ61 and AM60 salt spray surfaces as compared to the immersion surfaces. AE44 and AZ31 showed higher pit areas on the immersion surface as compared to the salt spray surface. The higher pit areas on the salt spray surfaces are due to the presence of pit debris covering the formed pits during the humidity and drying phases. Pit growth is considered autocatalytic, so once it starts, it continues unabated [1]. When pit debris covers the pit, as it does on the salt spray surfaces due to the inability of the humidity and drying phases to remove the pit debris, the pits can continue to grow without general corrosion interfering. When the pit debris is removed, either during the salt spray phase where water is present or during the cleaning process, the larger pits can be seen. While there is a minimal difference in pit area on the AZ31 surfaces, a difference is seen with the AE44 surface, though, because of the shift in corrosion location. When rare earth elements are present, corrosion shifts to the center of the grain and away from the intergranular region [7, 13]. This shift encourages general corrosion and pitting corrosion to "work together", thereby increasing the pit area on the immersion surface. In addition to the environment, though the form of magnesium plays a role in the pit area. The as-cast AM60 material has a smaller pit area on both environments as compared to the extruded AZ61 material. This again can be attributed to the as-cast skin on the AM60 material, which prevents pits from growing due to the small grain size. The extruded AZ61 material does not possess the as-cast skin, meaning that the pits can grow more easily. Again, AE44 does not follow this line, again likely due to the shift in pit formation location. Even though an as-cast skin exists on the AE44 material, the pits form within the magnesium grain. General corrosion works alongside pitting corrosion to degrade the magnesium grains, which grow together, indicating an increase in pit area. The magnesium alloy also appeared to play a role in the pit area, with the smallest pit area occurring on the ascast AM60 material, the largest pit area occurring on the as-cast AE44 material, and the middle pit area occurring on the extruded AZ61 and AZ31 materials. One could suspect that the presence of manganese affected the growth of the pits differently than the presence of either zinc or rare earth elements, but there is not currently enough alloys to accurately confirm this suspicion.
General corrosion, however, is only part of the model. The other portion of the model is pitting corrosion, which relates pit number density, pit area, nearest neighbor distance, and intergranular corrosion area fraction (ICAF). The first three values are highly interrelated. Pit number density and pit area are related because at the number of pits increase, the area covered by the pits increase. However, the pits can grow without the pit number increasing, meaning that pit area is not solely related to pit number density. Pit number density and nearest neighbor distance are also related, because as the pit increase in number, the distance between them decreases. Lastly, the pit area and nearest neighbor distance are related because as the pits grow in size, the distance between them decreases. These relationships are demonstrated in Figures 3-5. As the pit number density increases, so does the pit area, while the nearest neighbor distance decreases. The pit number density begins to decrease prior to pit area decreasing, as the pits can grow in size even when the pit number density decreases. In addition, the nearest neighbor distance decreases even as pit number density decreases, due to the slight growth in pit area. Once pit area begins to decrease, and pit number density continues to decrease, the nearest neighbor distance begins to increase. The differences seen in pit number density (Figure 3) can be attributed to the environment to which the magnesium is exposed, the form of magnesium, and the type of magnesium alloy. When it comes to the salt spray environment, fewer pits are seen on all surfaces as compared to the immersion surface. This is due to continuous exposure of chloride ions to the immersion surfaces, allowing the chloride ions to continually attack and pit the surfaces. However, chloride ions are only present on the salt spray surfaces for a limited time, meaning that the surfaces can
When it comes to the nearest neighbor distance (Figure 5), as previously mentioned, the pit number density and pit area greatly affect the nearest neighbor distance. The higher the pit number density and the larger the pits, the smaller the nearest neighbor
608
distance. When comparing the environments, since there are more pits formed on the immersion environments, the pits are closer together on the immersion surfaces. When comparing forms of magnesium, the as-cast AE44 surface and the extruded AZ61 surface had similar nearest neighbor distances because of the combination of pit area and pit number density, as did the ascast AM60 surface and the extruded AZ31 surface. Since the AZ61 surface had a higher pit number density but smaller pit areas as compared to the AE44 surface, the values combine to cause the nearest neighbor distance to appear similar. However, the as-cast AM60 magnesium alloy had the smallest pit number density and the smallest pit areas, the AM60 nearest neighbor distance would be the largest, or furthest apart. In addition, the extruded AZ31 magnesium alloy experienced a decrease in pit number density and an increase in pit area, meaning that the nearest neighbor distance was more influenced by the size of the pits instead of the number of pits. The pit area of AZ31, which was lOx higher than the other pit areas, ensured that the nearest neighbor distance started close and gradually increased, as the large pits incorporated other pits, increasing the distance between the remaining pits.
Conclusions Four magnesium alloys in two forms, as-cast AE44, as-cast AM60, extruded AZ61, and extruded AZ31 were examined in two corrosive environments, immersion and salt spray. General corrosion characteristics, weight loss and thickness loss, as well as surface characteristics, pit number density, pit area, nearest neighbor distance, and ICAF, were quantified over 60 hours. The most heavily corroded magnesium alloy, determined by combining general and pitting corrosion, was AZ61, followed by AE44, AZ31, and AM60, respectively. When comparing environments, more pits formed on all surfaces exposed to the immersion environment, while the pits were larger on the salt spray environments. References [1] M.G. Fontana, Corrosion Principles, in: M.G. Fontana (Eds.), Corrosion Engineering, McGraw-Hill, Boston, 1986, pp. 12-38. [2] G. Song, A. Atrens, "Understanding Magnesium Corrosion A Framework for Improved Alloy Performance", Advanced Engineering Materials 5 (2003) 837-858.
The intergranular corrosion area fraction (ICAF) is another measure of the coalescence of the pits, because as the pits that form along the intergranular boundary grow together, they eventually grew into each other, forming one long narrow pit. The type of magnesium alloy affected the ICAF much more than either the environment or the form of the magnesium alloy (Figure 6). When comparing the extruded AZ61 alloy with the as-cast AM60 alloy, once can see that the environment affected on the beginning of the ICAF, but by the end of the experiment time, the ICAF had merged between the salt spray and immersion environments. In addition, there was very little difference between the AM60 and AZ61 ICAF. However, there was a significant difference between AM60, AZ61, AZ31, and AE44. The difference is due to the presence of aluminum influencing the corrosion of AZ31 and the rare earth elements and the as-cast skin influencing the corrosion of AE44. For AZ31, there was 3% less aluminum in the magnesium alloy. Since aluminum, increasing to 10%, has been shown to increase the corrosion resistance of magnesium [2], it stands to reason that the lower percentage of aluminum in AZ31 would allow more intergranular corrosion. For AE44, both the alloying elements and the skin contributed to the formation of intergranular corrosion. Since rare earth elements switch corrosion from along the intergranular boundary to the interior of the magnesium grain [7,13] and the as-cast skin results in very small grains, the presence of ICAF means that the grains were degrading and connecting along the intergranular boundary. If there was no as-cast skin, meaning the grains were much larger, there is a chance that the ICAF would have been more in line with the AM60 and AZ61 samples. In addition to the differences caused by the presence of rare earth elements and the as-cast skin, the environment contributed to a difference in ICAF, with the immersion surface experiencing less intergranular corrosion than the salt spray surface. This difference can be contributed to general corrosion, which removed the intergranular boundaries that were left by the pitting and destruction of the magnesium grains. With general corrosion removing, or lowering, the intergranular boundaries, intergranular corrosion may not have been accurately quantified.
[3] BA Shaw, Corrosion Resistance of Magnesium Alloys, in: L.J. Korb, ASM (Eds.), ASM Handbook, Vol. 13A: Corrosion, Ninth Ed., ASM International Handbook Committee, Metals Park, 2003, pg. 692. [4] J.D. Majumdar, R. Galun, B. Mordike, I. Manna, "Effect of laser surface melting on corrosion and wear resistance of a commercial magnesium alloy", Materials Science and Engineering A, 361 (2003) 119-129. [5] C. Blawert, E.D. Morales, W. Dietzel, K.U. Kainer, "Comparison of Corrosion Properties of Squeeze Cast and Thixocast MgZnRE Alloys", Materials Science Forum, 488-489 (2005) 697-700. [6] W. Liu, F. Cao, L. Chang, Z. Zhang, J. Zhang, "Effect of rare earth element Ce and La on corrosion behavior of AM60 magnesium alloy", Corrosion Science, 51 (2009) 13341343. [7] W. Liu, F. Cao, L. Zhong, L. Zheng, B. Jia, Z. Zhang, J. Zhang, "Influence of rare earth element Ce and La addition on corrosion behavior of AZ91 magnesium alloy", Materials and Corrosion, 60 (2009) 795-803. [8] Y.L. Song, Y.H. Liu, S.R. Yu, X.Y. Zhu, S.H. Wang, "Effect of neodymium on microstructure and corrosion resistance of AZ91 magnesium alloy", Journal of Materials Science, 42 (2007) 4435-4440. [9] G. Song, "Recent Progress in Corrosion and Protection of Magnesium Alloys", Advanced Engineering Materials, 1 (2005) 563-586. [10] M.C. Zhao, M. Liu, G. Song, A. Atrens, "Influence of pH and chloride ion concentration on the corrosion of Mg alloy ZE41", Corrosion Science, 50 (2008) 1939-1953.
609
[11] G. Song, A. Atrens, X. Wu, B. Zhang, "Corrosion behavior of AZ21, AZ501, and AZ91 in sodium chloride", Corrosion Science, 40 (1998) 1769-1791. [12] Y.L. Song, Y.H. Liu, S.H. Wang, S.R. Yu, X.Y. Zhu, "Effect of cerium addition on microstructure and corrosion resistance of die cast AZ91 magnesium alloy", Materials and Corrosion, 58 (2007) 189-192. [13] N. Birbilis, M.A. Easton, A.D. Sudholz, S.M. Zhu, M.A. Gibson, "On the corrosion of binary magnesium-rare earch alloys", Corrosion Science, 51 (2009) 683-689. [14]
G. Song, A. Atrens, M. Dargusch, "Influence of microstructure on the corrosion of diecast AZ91D", Corrosion Science, 41 (1998) 249-273.
[15] S.P. White, G.J. Weir, N.J. Laycock, "Calculating chemical concentrations during the initiation of crevice corrosion", Corrosion Science, 42 (2000) 605-629. [16] R.M. Pidaparti, A. Puri, M.J. Palakal, A. Kashyap, "Twodimensional Corrosion Pit Initiation and Growth Simulation Model," Computers, Materials, and Continua, 2 (2005) 6575. [17] R.M. Pidaparti, L. Fang, M.J. Palakal, "Computational simulation of multi-pit corrosion process in materials", Computational Materials Science, 41 (2008) 255-265. [18] N.J. Laycock, J.S. Noh, S.P. White, D.P. Krouse, "Computer simulation of pitting potential measurements", Corrosion Science, 47 (2005) 3140-3177. [19] N.J. Laycock, J. Stewart, R.C. Newman, "The initiation of crevice corrosion in stainless steel", Corrosion Science, 39 (1997) 1791-1809. [20] L. Li, X. Li, C. Dong, Y. Huang, "Computational simulation of metastable pitting of stainless steel", Electrochimica Ada, 54 (2009) 6389-6395. [21] T. Johnsen, A. Jossang, T. Jossang, P. Meakin, "An experimental study of the quasi-two-dimensional corrosion of aluminum foils and a comparision with two-dimensional computer simulations", PhysicaA, 242 (1997) 356-276. [22] B. Malki, B. Baroux, "Computer simulation of the corrosion pit growth", Corrosion Science, 47 (2005) 171-182. [23] J.X. Jia, G. Song, A. Atrens, "Experimental Measurement and Computer Simulation of Galvanic Corrosion of Magnesium Coupled to Steel", Advanced Engineering Materials, 9 (2007) 65-74. [24] ASTM B117 - 07a (2007) Standard Practice for Operating Salt Spray (Fog) Apparatus, Vol. 03.02, 2007. [25] M.F. Horstemeyer, J. Lathrop, A.M. Gokhale, M. Dighe, "Modeling stress state dependent damage evolution in a cast Al-Si-Mg aluminum alloy", Theoretical and Applied Fracture Mechanics, 33 (2000) 31-47.
Magnesium Technology 2011 Edited by: Wim H. Sillekens, Sean R. Agnew, Neale R. Neelameggham, andSuveen N. Mathaudhu TMS (The Minerals, Metals & Materials Society), 2011
CORROSION PERFORMANCE OF Mg-Ti ALLOYS SYNTHESIZED BY MAGNETRON SPUTTERING Zhenqing Xu1, Guang-Ling Song2*, Daad Haddad1 'MEDA Engineering and Technical Services, LLC, 17515 W 9 Mile Rd, STE 1075, Southfield, MI 48075, USA 2 General Motors Research and Development, Mail Code: 480-106-224, 30500 Mound Road, Warren, MI 48090, USA Keywords: Corrosion, Mg, Ti, Magnetron Sputtering, XPS Abstract
Experimental Film Deposition
Mg is difficult to alloy with Ti through a conventional metallurgical approach due to their insolubility in each other and big difference in melting point. However, Mg, if alloyed with Ti, may become corrosion resistant. This hypothesis is verified in this study.
Mgi-xTix (x=0, 0.2, 0.4, 0.6, 0.8, 1) thin films were deposited onto a round glass disk, 2 cm in diameter, using dc magnetron sputtering in an argon atmosphere at room temperature. The base pressure in the growth chamber was about 5xl0" 8 Torr. The dynamic pressure during film deposition was 2 mTorr. High purity Mg and Ti ingots were used as targets. The thin film was very uniformly formed on the glass substrate. Film composition was controlled by changing the power applied to the Mg and Ti targets. The compositions of the different Mg-Ti thin film samples were confirmed by Electron Probe Microanalysis (EPMA).
Mg,.xTix alloy thin films (with x=0, 0.2, 0.4, 0.6, 0.8 and 1) were deposited by magnetron sputtering onto a glass substrate. Film compositions were analyzed by electron probe micro-analysis (EPMA). The electrochemical behavior of these alloys was characterized in saturated Mg(OH)2 solutions with and without 0.1 M NaCl. The macrostructures of the thin film alloys were compared before and after polarization and immersion measurements. The results showed that the corrosion resistance of the alloy was improved with increasing Ti content. No material loss or corrosion damage was observed for alloys with 80% or more Ti content in both solutions.
Electrochemical Characterization A glass flat cell and a Solatron 1280 potentiostat system were used for polarization curve measurements. The sample was immersed in the solution at its open circuit potential (OCP) for 5 min before polarization curve measurement. The immersed area was 1 cm2. A platinum gauze (2.5 cm x 2.5 cm) was used as the counter electrode and a KCl-saturated Ag/AgCl electrode was used as the reference (Ref). Potentiodynamic polarization curve was recorded at a potential scanning rate of 0.1 mV/s from -0.2 V vs. OCP to+1.0 V vs. Ref.
Introduction Mg alloys have found many potential applications recently. Their low density and high strength to weight ratio make Mg alloys attractive to the auto industry in weight reduction and efficiency improvement. However, many of the alloying elements form intermetallic phases with Mg which have a more positive free corrosion potential than the Mg matrix itself [1-3]. Hence, microgalvanic corrosion attacks usually decrease the overall corrosion resistance of a Mg alloy. It is hypothesized that a precipitation free, single phase, solid solution magnesium alloy will have better corrosion performance.
Immersion Test and Surface Film Characterization Mg-Ti thin film samples were immersed in Mg(OH)2+0.1M NaCl solution for 4 days. Their morphologies after immersion were examined under optical microscope and Scanning Electron Microscope (SEM). A small section in the immersion area was analyzed by X-ray photoelectron spectroscopy (XPS). High resolution XPS has also been performed on Mg 2p, O Is, and Ti 2p core levels. As is the standard practice in XPS studies, the Cls line corresponding to the C-C bond has been used as the binding energy (BE) reference [9].
Mg and Mg alloys present a very negative open circuit potential in a sodium chloride solution and have an active and fast anodic dissolution process [4]. The addition of strong passivation elements like Cr [5] and Ti [6-8] could possibly improve the passivity of a Mg alloy. However, according to the phase diagram, the solubility of Mg in Ti, and vice versa, is very small in crystalline state. Conventional equilibrium metallurgical approach cannot successfully lead to a Mg-Ti solid solution phase due to low solid solubility in each other and very high Ti melting point relative to Mg. Using a non-equilibrium magnetron sputtering deposition technique can substantially extend the solubility of Ti in the Mg-Ti alloy and in this case the large melting point difference between Ti and Mg is no longer an issue.
Results and Discussion Deposition and Film Composition The deposition conditions for growing Mg-Ti thin films and the EPMA composition characterization are listed in Table I. By changing the power applied to the Mg and Ti targets, different growth rates of Mg and Ti can be achieved, forming Mg-Ti thin films with different Ti contents. The EPMA measurements confirm that the deposition parameters used in sputtering quite accurately controlled the thin film compositions. Thickness of these thin films has been measured by cross section SEM. These samples have thickness ranging from 1.2 |im to 1.5 p.m.
In this paper, Mg-Ti thin film alloys with different Ti concentrations were deposited and their corrosion behavior was compared, aiming to understand the effect of the Ti solute on the corrosion behavior of Mg.
611
Table I. EPMA composition characterization of Mg-Ti thin films deposited by magnetron sputtering Power (W)
Sample
Rate (À/s)
Calculated Concentration (Using Growth rates)%at
Measured Concentration (EPMA)%at
Measured Thickness (Cross section SEM) um
Mg
Ti
Mg
Ti
Mg
Ti
Mg
Ti
Mg
100
-
4.6
-
100
-
>99
<1
1.2
Mg80Ti20
100
130
4.6
0.9
80
20
79
21
1.3
Mg60Ti40
50
200
2.5
1.3
60
40
59
41
1.3
Mg40Ti60
40
310
2.0
2.1
40
60
42
58
1.5
Mg20Ti80
20
450
1.0
2.9
20
80
19
81
1.4
Ti
-
300
-
2.0
-
100
<1
>99
1.3
Mg80Ti20. Crevice corrosion was found at the areas that were in contact with the plastic washer of the electrolyte cell. The severity of the crevice corrosion is in reverse order of the Ti content in the Mg-Ti alloys. The above macroscopic observations show that with increased Ti content, a Mg-Ti alloy thin film becomes more resistant to general corrosion during polarization. Their ability to preserve more material when being anodized with high overvoltage has increased significantly. It is widely known that the corrosion of Ti results in a stable oxide layer (essentially Ti02) on the surface. The improved passivity of a Mg-Ti alloy with increasing Ti content could result from formation of more Ti0 2 on the surface.
Corrosion Behavior To investigate the effect of Ti on Mg-Ti alloy's electrochemical behavior, potential-dynamic polarization curves were measured in Mg(OH)2 solutions, with and without 0.1 M NaCl. Polarization curves of sputter-deposited pure Mg and Ti were also measured for comparison. Figure 1 shows macroscopic images of the Mg-Ti alloys after polarization in both solutions. The top part of the coupon sample has undergone the polarization measurements in the Mg(OH)2 saturated solution, while the bottom part is in the solution with 0.1M sodium chloride addition. For samples polarized in saturated Mg(OH)2, significant material loss can be observed for pure Mg, Mg80Ti20 and Mg60Ti40. Particularly for pure Mg, no material was preserved after polarization to +1 V vs. Ref. Obviously, a film with increased Ti content was less corroded, indicating that a increase in Ti content can lead to improved corrosion resistance.
Figure 2 shows the polarization curves of Mg-Ti alloys measured in saturated Mg(OH)2 without (a) and with (b) 0.1 M NaCl. The influence of the noble Ti on the free corrosion potential can be clearly identified. The higher the Ti content, the more positive is the corrosion potential. In Mg(OH)2 solution shown in Figure 2a, very little passive tendency is observed for Mg and Mg80Ti20. As the polarization potential is more positive than +0.5 V vs. Ref the current density deceases significantly. This phenomenon cannot be linked to passivation of the material. Rather, considering the previous observation from Figure 1, we can conclude that at a voltage more positive than +0.5 V vs. Ref, significant material was corroded for these two samples and the inert substrate glass was exposed to the solution, resulting in decreased anodic dissolution current densities. A clear passive regime with constant current can be seen in the anodic region for samples with Ti content larger than 40%. Although Mg60Ti40 has the lowest passivation current density, it loses its passivation when potential is more positive than 0.25 V vs. Ref. Other than Mg60Ti40, the passivation current densities decrease with more Ti addition into Mg-Ti for Mg-Ti alloy when Ti content is higher than 60at%. This higher corrosion resistance can be attributed to structure modification of Mg film by substitution of Mg with Ti and improved stability of the surface film due to Ti incorporation.
Figure 1. Macroscopic images of Mg-Ti thin films on glass disks after polarization in saturated Mg(OH)2 (top round areas) and Mg(OH)2+0.1 M NaCl (bottom round areas). With the addition of 0.1 M sodium chloride, similar behaviors of these films were observed. Apparent corrosion can be found for all samples other than pure Ti and Mg20Ti80, although much of the materials were preserved for Mg40Ti60, Mg60Ti40 and
612
material loss or cracking can be observed for Mg80Ti20; only a slight color change is found. It was previously shown in Figure 2 that under polarization Mg60Ti40 was more corrosion resistant than Mg80Ti20. However, an interesting finding is that after 1 hour immersion Mg60Ti40 experienced more material loss from the substrate than Mg80Ti20. For Mg-Ti alloys with over 40 at.%Ti, no corrosion occurred on their surface after immersion. After 4 hour immersion (Figure 3b), all Mg was corroded away for pure Mg thin film. There is no significant material loss for Mg80Ti20. A gap between the top and bottom parts is found, suggesting that some of the film was dissolved along the waterline. Mg60Ti40 has lost considerable amount of the film after 4 hours of immersion. No corrosion can be detected on all other samples.
Figure 2. Polarization curves of Mg-Ti thin films in (a) in saturated Mg(OH)2 and (b) saturated Mg(OH)2+0.1 M NaCl; Similar electrochemical behavior has been observed in solution of Mg(OH)2+0.1M NaCl (Figure 2b). Because of the presence of corrosive Cl" in the solution, no passivity can be detected for pure Mg and Mg80Ti20. Both Mg60Ti40 and Mg40Ti60 have a small passive region while the latter one has a higher passivation potential and a broader passive potential region from -0.6 to -0.2 V vs. Ref. A dramatic anodic dissolution current density decrease was recorded for Mg-Ti thin films with Ti content less than 40%. A possible scenario is that most of the material was corroded at the crevice corrosion region (shown as a corroded ring in Figure 1) and became electrically insulating, leading to a diminished current density shown in Figure 2b. Mg20Ti80 and pure Ti stay in the passive region even when the polarization potential is at +1 V vs. Ref. Immersion Test Mg-Ti thin films morphologies were recorded using an optical microscope for different immersion times in the saturated Mg(OH)2+0.1M NaCl solution (Figure 3). The top part of each image is where the sample was immersed. It was found that after 1 hour of immersion pure Mg was corroded significantly with pitting and cracking evident all over the immersed area. No
Figure 3. Macroscopic images of Ti-Mg thin film alloys after immersion in Mg(OH)2+0.1M NaCl for (a) Ihr and (b) 4 hr. The top part of the graph is the immersed area. Further analysis of the structure of these Mg-Ti thin films after immersion test is presented by SEM in Figure 4. Under high magnification cracks can be found in Mg80Ti20, Mg60Ti40 and Mg40Ti60 samples after 4 hour immersion. It seems that the most severe film rupture and cracking is found in Mg60Ti40 (Figure 4b). The density of micro-cracks in Mg80Ti20 is the highest; however, the film seems to adhere very well to the glass substrate. There is no film peeling or delamination. The SEM observation
spectra revealed the presence of Mg 2p, O Is and Ti 2p. These characteristic peaks correspond to the electron configuration of the electrons within the atoms. High resolution Mg 2p XPS spectra (Figure 5a) show that one broad peak at 50.8 eV is presented for both specimens, suggesting the existence of MgO and Mg(OH)2. The O Is spectra presented in Figure 5b show different characteristics for these two samples. The Mg20Ti80 sample has only one main O Is peak presented at 531.0 eV, corresponding to oxide. The O Is composite peak of Mg80Ti20 sample can be deconvoluted to two big peaks, corresponding to oxide (531.0 eV) and hydroxide (532.8 eV). The measured high resolution peaks for Ti 2p are shown in Figure 5c. The peak positions 459.0 for Ti oxide is in very good agreement with the literature values given for Ti02 (458.7 eV) [7].
confirms the previous macroscopic finding that Mg60Ti40 has more material loss than Mg80Ti20 after 4 hours of immersion test. Although polarization results suggest that Mg60Ti40 shows more passive nature than Mg80Ti20 (Figure 2), more concentrated stress presented in the film could lead to peeling and delamination after the sample is taken out from the solution and exposed to air. Further analysis such as X-ray diffraction (XRD) and atomic force microscopy (AFM) are being conducted to investigate why stress is more uniformly presented in Mg80Ti20. The grain size, orientation or mechanical properties of the Mg-Ti thin films may all play a role in the stress distribution. Moreover, the difference in strength and brittleness of the compounds could also results in different film cracking patterns.
(a)Mg2p i i
■ il*
h
i
u
Mg20Ti80
*
Mg80Ti20 54
52
50
46
48
44
Binding Energy (eV)
(b) O Is
Mg20Ti80 Figure 4. SEM micrographs of Ti-Mg thin film alloys after immersion in Mg(OH)2+0.1 M NaCl.
Mg80Ti20 538
Immersion test was continued for 4 days for Mg40Ti60, Mg20Ti80 and pure Ti. There is still no cracking or material loss observed for pure Ti and Mg20Ti80 (Figure 4e and f). A small amount of Mg40Ti60 was still left on the substrate (Figure 4d) while Mg80Ti20 and Mg60Ti20 have lost all of the deposited films. Overall, the improvement of corrosion resistance is evident when the Ti content is increased. This beneficial effect observed for Mg-Ti alloys could be partially attributed to the excellent corrosion resistance of Ti. Corrosion of Mg occurs at breaks in a partially protected surface film. The passive film Mg(OH)2 is mechanically weak and unstable and not able to prevent further corrosion of Mg. The formation of continuous Ti0 2 passive layer should play an important role in increasing the corrosion resistance of Mg-Ti alloys.
536
534
532
530
528
526
Binding Energy (eV)
(c)Ti2p
Mg20Ti80 Mg80Ti20 480
475
470
465
460
455
450
445
Binding Energy (eV)
XPS Characterization XPS measurement was conducted on two selected samples, Mg20Ti80 and Mg80Ti20 after 4 hour immersion test. The survey
Figure 5. XPS high resolution spectra of Ti-Mg after immersion in Mg(OH)2+0.1MNaCl.
614
Cr implanted Mg surfaces, Surface & Coatings Technology, 158 (2002) 328-333.
The above findings suggest that the Mg20Ti80 sample surface has a strong oxide layer, possibly Ti0 2 . It confirms our previous speculation that a stable oxide layer could form on the alloy with higher Ti content and prevent corrosion attack. A small peak is also presented for Mg20Ti80 sample at 453.0, corresponding to Ti in its metallic state. For the Mg80Ti20 sample, Mg is not only presented in its oxide state, but also in the surface film Mg(OH)2. Unfortunately, this film is not very protective [10, 11]. Only a higher concentration of Ti in the film can beneficially enhance the passivity. The mixture of Ti0 2 and MgO forms a compact passive film preventing the attack of aggressive Cl" ions, leading to improved corrosion resistance of a Mg-Ti alloy with a higher Ti content.
[6] K.R. Baldwin, D.J. Bray, G.D. Howard, R.W. Gardiner, Corrosion behaviour of some vapour deposited magnesium alloys, Materials Science and Technology, 12 (1996) 937-943. [7] T. Mitchell, S. Diplas, P. Tsakiropoulos, Characterisation of corrosion products formed on PVD in situ mechanically worked Mg-Ti alloys, Journal of Alloys and Compounds, 392 (2005) 127141. [8] S. Rousselot, M.P. Bichat, D. Guay, L. Roue, Structure and electrochemical behaviour of metastable Mg50Ti50 alloy prepared by ball milling, Journal of Power Sources, 175 (2008) 621-624.
Summary The corrosion properties of Mg x Ti]. x thin films with x ranging from 0 to 1 were investigated. Incorporation Ti in the Mg lattice has been achieved by dc magnetron sputtering. Polarization measurements showed that alloys with high Ti content had better passivity in solutions of Mg(OH)2 with and without 0.1M NaCl. XPS observation confirms the presence of surface oxide film on the Mg-Ti alloy surface. A high ratio of oxide/hydroxide in the Mg20Ti80 sample surface is responsible for improved corrosion resistance.
[9] M. Liu, S. Zanna, H. Ardelean, I. Frateur, P. Schmutz, G. Song, A. Atrens, P. Marcus, A first quantitative XPS study of the surface films formed, by exposure to water, on Mg and on the Mg-Al intermetallics: A13Mg2 and Mgl7A112, Corrosion Science, 51(2009)1115-1127.
Production of supersaturated single-phase compounds of Mg-Ti coatings by sputtering provides a promising approach to produce corrosion resistance magnesium based alloys outside the conventional casting technology. Formation of precipitates during sputtering is strongly reduced. As a result, high purity coatings can be obtained and a reduced corrosion rate can be achieved. Compound composition strongly influences the properties and corrosion behaviors of the alloy films. Therefore, other corrosion resistant materials could also be used to form alloys with Mg to prevent corrosion.
[11] A. Seyeux, M. Liu, P. Schmutz, G. Song, A. Atrens, P. Marcus, ToF-SIMS depth profile of the surface film on pure magnesium formed by immersion in pure water and the identification of magnesium hydride, Corrosion Science, 51 (2009) 1883-1886.
[10] G. Song, Recent progress in corrosion and protection of magnesium alloys, Advanced Engineering Materials, 7 (2005) 563-586.
Acknowledgements The authors would like to thank Mr. Richard Waldo for EPMA measurements and Mr. Nicholas Irish for XPS analysis. References [1] A. Pardo, M.C. Merino, A.E. Coy, R. Arrabal, F. Viejo, E. Matykina, Corrosion behaviour of magnesium/aluminium alloys in 3.5 wt.% NaCl, Corrosion Science, 50 (2008) 823-834. [2] G. Song, Investigation on corrosion of magnesium and its alloys, Journal of Corrosion Science and Engineering, 6 (2003). [3] G. Song, B. Johannesson, S. Hapugoda, D. Stlohn, Galvanic corrosion of magnesium alloy AZ91D in contact with an aluminium alloy, steel and zinc, Corrosion Science, 46 (2004) 955-977. [4] G. Song, Z. Xu, The surface, microstructure and corrosion of magnesium alloy AZ31 sheet, Electrochimica Acta, 55 (2010) 4148-4161. [5] M. Vilarigues, L.C. Alves, I.D. Nogueira, N. Franco, A.D. Sequeira, R.C. da Silva, Characterisation of corrosion products in
615
Magnesium Technology 2011 Edited by: Wim H. Sillekens, Sean R. Agnew, Neale R. Neelameggham, andSuveen N. Mathaudhu TMS (The Minerals, Metals & Materials Society), 2011
STRUCTURE AND MECHANICAL PROPERTIES OF MAGNESIUM-TITANIUM SOLID SOLUTION THIN FILM ALLOYS PREPARED BY MAGNETRON-SPUTTER DEPOSITION Daad Haddad1, GuangLing Song2*, Yang Tse Cheng3 'MEDA Engineering and Technical Services LLC; 17515 W. Nine Mile Road; Southfield, MI 48075, USA 2 Chemical Sciences and Materials Systems Lab, GM Global Research and Development; 30500 Mound Road; Warren, MI 48090, USA department of Chemical and Materials Engineering, University of Kentucky; 177 F. Paul Anderson Tower; Lexington, KY 40506, USA Keywords: Magnetron Sputtering, Mg-Ti thin films, Nanoindentation, X-ray diffraction, Atomic Force Microscopy structure of these films are studied using AFM and XRD measurements. Finally, the mechanical properties of these films are measured using nanoindentation experiments and by analyzing the load-displacement curves based on the Oliver-Pharr method [16]. The effect of the composition and morphology on the mechanical properties of these alloys is discussed.
Abstract Mg alloys are being considered for wider application in automotive industry. Designing new alloys with improved mechanical properties is important to the development of new Mg alloy parts. Mg-Ti is an interesting alloying system that may have good corrosion resistance due to high passivity of Ti. However it is difficult to form through a conventional metallurgical method due to the mutual insolubility of Mg and Ti and the big difference in their melting point. Nevertheless, if the alloy can be formed, it may have other unexpected physical and chemical performance. Therefore, it is of significance to understand the properties of MgTi alloy produced by non-conventional approach.
Mg-Ti Thin Films Deposition
In this report, Mg^.^Ti,, thin film alloys containing 0, 21, 41, 51, 58, 81 and 100 at.% Ti were deposited by dc magnetron sputtering on Si substrates. The mechanical properties of the thin film alloys were obtained using nanoindentation. Electron probe microanalysis (EPMA) was used to determine the film compositions. X-ray diffraction (XRD) measurements showed that single phase magnesium-titanium solid solutions were obtained across the full range of magnesium and titanium mixtures. The topography and the rms roughness of the different alloys were studied using atomic force microscopy (AFM). The mechanical properties of the Mgd.x)Tix thin films were determined by analyzing the nanoindentation load-displacement curves based on the Oliver-Pharr method. The nanoindentation results show that both the elastic modulus and hardness of the Mg(i_X)Tix alloy thin films are higher than those of conventional Mg alloys. Introduction Magnesium and magnesium alloys have attracted wide attention as lightweight materials that can be used in automotive, aerospace, and hydrogen storage applications. The alloying of magnesium and titanium is of particular interest due to the improved mechanical and corrosion resistant properties of Ti [1-5]. However, magnesium and titanium have very little mutual solubility according to their equilibrium phase diagram, and thus do not form any stable intermetallic compounds under standard alloying conditions [6, 7]. But recently it was reported that by using non-standard alloying techniques, such as electron-beam deposition and magnetron sputtering, the solid solubility of Mg and Ti can be extended significantly and solid solution Mgfl.x)Tix thin film alloys can be formed [8-15]. These Mg(i_x)Tix films are expected to have the advantages of being lightweight alloys with high specific strength and better corrosion performance.
The Mg,,.x)Tix (x = 0, 0.21, 0.41, 0.51, 0.58, 0.81, 1) thin films were grown on 1" Si (100) substrates using dc magnetron sputtering under argon atmosphere. The silicon substrates were ultrasonically cleaned in acetone and methanol successively for 20 minutes. The cleaned substrates were loaded in the load lock chamber where they were heated to 200°C for 30 minutes under vacuum for outgasing. After cooling down, the substrates were transferred to the growth chamber which has a base pressure of about 5. Ox 10"8 Torr. The deposition plasma for each constituent material was created utilizing a dc power source and a flow of Ar (14 seem). The dynamic pressure during the growth of the films was 2 mTorr. Deposition of the different alloys is carried out using a recipe which depends on predetermined deposition rates. Growth rates of the magnesium and titanium sources were measured, under the above specified growth conditions, at different dc powers (50 600 W) using deposition monitor with the crystal sensor positioned just above the center of the substrate holder. For each composition, the ratio of Mg to Ti growth rate was calculated as a function of the ratio of Mg to Ti atoms. Using the measured growth rates and the calculated growth rate ratio, the film composition was controlled by changing the power to the Mg and Ti targets. The dc powers applied to each target during the deposition of the different alloys together with EPMA determined compositions are summarized in Table I. There is excellent agreement between the measured and calculated compositions. The substrates were rotated at 20 rpm in order to obtain uniform composition across the substrates. The films are typically 1.2-1.5 um thick as determined from SEM cross section measurements. Materials Characterization The composition of the different Mg-Ti thin film samples was determined using Electron probe microanalysis (EPMA). EPMA measurements were made with a Cameca Instruments, Inc. (Nampa, ID) model SX-100 electron probe. Analysis voltage and current were 15 keV and 10 nA. Analyses were done with a focused beam < 1.0 um in diameter. Each sample was analyzed at 8 random locations with results averaged. Mg and Ti compositions were calculated from x-ray intensities (k-ratios) with the thin film
In this work, the fabrication of Mg(i_x)Tix thin film alloys with different Ti content is presented. The topography and the crystal
617
program GMRFILM [17]. Precision and accuracy are both estimated as +/- 2% relative for all samples.
Results and Discussion Crvstallographic Structure
Ex-situ X-ray diffraction was used to study the structures of the Mg-Ti films. Each sample was examined using Cu K„ radiation in a Bruker AXS General Area Detector Diffractometer System (GADDS). The diffraction images are typically collected for a period of 5 minutes using a 0.5 mm collimator and a sample to detector distance of 60 mm. The primary beam incidence angle is 17 degrees. Each diffraction scan was examined for incomplete Debye rings, which indicate preferred crystallite orientation, crystallite size in the film, and single crystal substrate diffraction peaks. Each diffraction image is then integrated to produce a powder diffraction pattern. The diffraction patterns are then compared to reference data for phase identification. Once phase identification is complete lattice parameters can be calculated.
The microstructure of the Mgd_x)Tix thin film alloys was determined by XRD measurements. The XRD patterns of the different Mg-Ti films after normalization are shown in Figure 1. The Mg film shows a very strong diffraction peak at 26 ~ 34.8° which is due to the (0002) plane of hexagonal closed packed (HCP) Mg. The Mg film exhibits a strong preferential crystallographic orientation with the (0001) plane growing parallel to the substrate surface. The weak diffraction peaks at higher 26 correspond to the HCP Mg (10Î2) and (10Ï3) peaks. XRD pattern of the Mgo.79Tio.21 thin film shows a strong diffraction peak at 26 ~ 35.8° that is due to the (0002) plane of HCP solid solution unit cell. This peak is shifted to a higher angle (lower d-spacing) with respect to pure Mg due to the partial substitution of Mg atoms by Ti atoms. Because Ti has relatively smaller molar volume than Mg, partial substitution of Mg by Ti causes the Mg lattice to contract and thus d-spacing to decrease. The peak at higher 26 is due to (10Ï3) HCP Mg-Ti solid solution. Mg and Mgo.79Tio.21 samples have a strong [001] fiber texture as reflected by the strong (0002) diffraction peak of the HCP Mg-Ti unit cell.
Topography measurements of the Mg-Ti alloy samples were obtained using Dimension V Scanning Probe Microscope (Veeco). The height measurements were obtained using tapping mode and scanning area of 5.0 um x 5.0 um. The AFM measurements were performed in air, at ambient temperature and humidity, using a silicon probe with tip ROC (radius of curvature) < 10 nm. Four topography maps at different positions were acquired for each sample. The reported values of the rms roughness are the mean values of the different measurements.
TirtXXH)
Nanoindentation experiments were carried out with a Hysitron Triboscope (Hysitron Incorporated, Minneapolis, MN). The load controlled indents were made with a Berkovich indenter. The load function consists of three segments, a 5 seconds loading segment, a 2 seconds holding segment at the maximum load, and a 5 seconds unloading segment. Indents were made to a depth of about 10 percent of the film thickness to minimize substrate effects (typically -140 nm). Data values are averaged from a series of 16 indents per sample. The hardness and modulus are calculated based on the Oliver-Pharr method [16].
Ti (10ïft) jlyyTi (1011)
MaTi(1
Table I. The dc power applied to the Mg and Ti sources during deposition of Mgd.x)Tix thin films at room temperature along with the corresponding Mg and Ti concentration calculated using measured growth rates and the concentration determined from EPMA measurements
Sample
Calculated Concentration (Using Growth rates) (at.%)
Power (W)
Ti
Mg
Ti
Mg
Ti
Mg
100
-
100
-
>99
<1
Mgo,8oTio.2o
100
130
80
20
79
21
Mgo.6oTio.4o
50
200
60
40
59
41
Mgo.5oTio.50
50
300
50
50
49
51
Mgo.4oTio.6o
40
310
40
60
42
58
Mgo.2oTio.8o
20
450
20
80
19
81
Ti
_
300
_
100
<1
>99
MgTi^m
i . . . .
25
■ , , . .
Ti
Mgoi9
AM9(0002)Mg.(1OT2) Mg(10Î3)
Measured Concentration (EPMA) (at.%)
Mg
Ti(112Q) TK10Ï3)
^
Mg
i
30 35 40 45 50 55 60 65 70 75 80 29 (deg.)
Figure 1. XRD spectra of Mg(i.x)Tix thin film alloys. The film with 41 at.% Ti shows a single phase corresponding to HCP Mg-Ti solid solution. XRD pattern of this film shows three diffraction peaks at lower 26 (~ 34.0°, 36.6°, and 38.7°) which are due to the (10Ï0), (0002), and (10Î1) peaks of a contracted HCP Mg lattice with Mg atoms substituted by Ti. The Mg(i.X)Tix samples with x ~ 0.51, 0.58 and 0.81 show also a single phase with two diffraction peaks that can be attributed to the(10T0), and (10Ï1) diffraction peaks in HCP Mg-Ti lattice. In all these films, the higher 26 diffraction peaks correspond to the Mg-Ti HCP phase. The pure Ti film, unlike the Mg film, does not show one highly preferred orientation. The XRD spectra of Ti film show the (10Î0),(0002), and (10Ï1) diffraction peaks of Ti HCP lattice at lower 26 (35.5°, 38.6°, and 40.4 °).
618
The above XRD results suggest that all the Mg-Ti thin film alloys are a single hexagonal Mg matrix phase with various contents of Ti in the solid solution. The lattice of the Mg-Ti solid solution contracts with the addition of Ti. In Figure 2, the lattice constants of the Mg(i.x)Tix films calculated from XRD spectra are presented as a function of Ti content, x. Also shown in Figure 2 are the lattice constants at different composition calculated using Vegard's Law [18]. Accordingly, the lattice constants of the alloys are linearly interpolated from the values of the pure compounds using the following relation: %I-*>B*
=
* 1 -*)aA+
xaB
maximum value for the sample with 58 at.% Ti (Figure 4). The topography of the Ti-rich sample Mgo.19Tio.8i shows long narrow rectangular stick-like particles. These particles grow evenly along the surface of the film and do not protrude out of the surface making the surface of this film smooth with small surface roughness compared to the films with 51 and 58 at.% Ti.
(1)
The calculated lattice constants from XRD spectra agree with those predicted using Vegard's law confirming that a solid solution of Mg-Ti was obtained.
Figure 2. Dependence of lattice constants on the composition of the Mg-Ti samples. Also shown the linear dependence of lattice constants on the alloy composition predicted by Vegard's Law. Topographic morphology AFM height images of the Mg-Ti alloys are given in Figure 3. The pure Mg sample shows large hexagonal particles protruding from the film surface at different angles, and making the sample surface very rough. Figure 4 displays the root-mean-square (rms) roughness of the Mg-Ti thin films obtained from AFM height images as a function of Ti concentration. These results show that Mg thin film has a very high rms roughness of ~ 30nm. The AFM image of sample Mgo.79Tio.21 shows that the large hexagonal particles disappeared and are replaced by small irregularly shaped particles that stretch along the sample surface. The refined almostflat particles in sample Mgo 79Ti0 21 leave the surface of this sample significantly smoother with a much smaller rms roughness compared to Mg sample (Figure 4). As the Ti concentration increases to 41, 51, and 58 at.%, (1010) and (10T1) oriented Mg-Ti grains start to emerge in these films. AFM height images of these samples show also angular particles that stick out of the film surface. The appearance of these new (10T0) and (10T1) oriented Mg-Ti grains in these films affects the surface roughness of these films, which increases to a
Figure 3. AFM topography images of the different Mg-Ti alloy thin films: a) Mg, b) Mgo.79Tio.21, c) Mgo.59Tio.41, d) Mgo.49Tio.51, e) Mgo.42Tio.j8, 0 Mgo.19Tio.8i, and g) Ti.
619
Finally, the pure Ti topography shows the fine rectangular sticklike shape as in Mgo i9Ti0 81 but in smaller size in addition to some coarse larger particles making the surface roughness of this film slightly higher than that of the Mgo i9Ti08i sample
geometry and S is the stiffness of the unloading curve given by the slope of the initial portion of the unloading curve. A total of 16 indents are performed on each sample and the load was chosen so that the contact depth was ~ 140 nm to minimize substrate and roughness effects. The calculated mechanical properties of each sample are then averaged and the mean values of modulus and hardness are plotted in Figure 5 along with the standard deviations. The measured Young's modulus and hardness of pure Mg and Ti thin film samples clearly show, as expected, higher Young's modulus and hardness values of Ti sample over Mg sample. It is reasonable to expect the Young's modulus and hardness of the Mg-Ti alloys to increase monotonically with the addition of Ti. However, the results show that Mg(i_x)Tix samples with x = 0.21, and 0.81 have both similar Young's modulus and hardness. Actually, sample Mgo79Tio2i shows very large improvement in mechanical properties compared to pure Mg sample. But, as Ti content increases to 41, 51, and 58 at.% the measured Young's modulus and hardness of the corresponding samples are lower than those of the Mg-Ti sample with 21 at.% Ti but higher than those of pure Mg sample.
The in-plane particle sizes of the different Mg-Ti samples were calculated using cross section analysis of the AFM height measurements and are presented in Figure 4. For each sample the measured lengths of 15 different particles along the short and long axes were averaged to give a mean value of the particle size ofthat sample. Mg sample shows large particle size of about ~ (750 ± 120) nm. The particle size value decreases to ~ 200 nm for the samples with Ti content of 21, 41 and 51 at.%, and then slightly increases to ~ (250 ± 80) nm for the Mgo42Tio58 sample. For the Ti-rich sample (x = 0.81) the particle size decreases again to ~ (175 ± 90) nm reaching a minimum value of- (140 ± 40) nm for the pure Ti sample.
Figure 4. Root-mean-square (rms) roughness and particle size of Mg-Ti alloy thin films, obtained from AFM height images and displayed as a function of Ti concentration. Micro-Mechanical Properties
Figure 5. Calculated Young's modulus and hardness of the Mg-Ti alloy samples plotted as a function of Ti content.
The commonly used Oliver-Pharr method was used to analyze the indentation data and to calculate the Young's modulus and hardness of the Mg-Ti samples. According to this method, the hardness is defined as: "
—
Fmax/"c
These results can be understood once the effect of roughness on the measured modulus and hardness values is taken into account [19, 20]. This effect originates from the initial contact between the indenter tip and the rough surface when some flattening can occur causing the measured hc and thus A,, to be less accurate. As a result the measured modulus and hardness of the Mg-Ti samples are affected by both the addition of Ti content which should result in improving mechanical properties and the increased surface roughness of the samples which would reduce the values of the modulus and hardness. The Mgo79ÏÏ0.21 shows a very small surface roughness compared to Mg sample, and as a result its mechanical properties are strongly improved due to both Ti addition and lower surface roughness. The surface roughness of Mg-Ti samples with 41, 51, and 58 at.% Ti increases and reaches a maximum value for the sample with 58 at.% Ti. As a result, the modulus and hardness of these samples are lower than the sample with 21 at.% Ti even though they are more Ti-rich. As Ti content increases to 81 at.%, the surface roughness of the Mgo.19Tio.8i decreases. The effect of increased Ti content on the mechanical properties of this film is
(2)
where Pmax is the maximum indentation load and A,, is the projected contact area of the indenter tip at that load. The contact area of the indenter is calculated as a function of contact depth, hc, using a series of indents at various loads (various hc) performed on a sample with a known elastic modulus. Although this method does not account for the resulting indentation shape (e.g., pile-up) or the elastic mismatch between the film and the substrate it serves as a tool for estimating mechanical properties. The Young's modulus is obtained from:
EB*/YWS/i)(S/JÄd
(3)
where y is a correction factor that depends on the indenter
620
9. P. Vermeiden, H. J. Wondergem, P. C. J. Graat, D. M. Borsa, H. Schreuders, B. Dam, R. Griessen and P. H. L. Notten, "In situ electrochemical XRD study of (de)hydrogenation of MgyTi10o-y thin films," Journal of Materials Chemistry, 18 (2008), 36803687.
apparent by the increased modulus and hardness values. Conclusion Thin films of Mg
10. B. Farangjs, P. Nachimuthu, T. J. Richardson, J. L. Slack, B. K. Meyer, R. C. C. Perera, and M. D. Rubin, "Structural and electronic properties of magnesium-3D Transition metal switchable mirrors," Solid State Ionics, 165 (2003), 309-314. 11. D. M. Borsa, A. Baldi, M. Pasturel, H. Schreuders, B. Dam, R. Griessen, P. Vermeulen, and P. H. L. Notten, "Mg-Ti-H thin films for smart solar collectors," Applied Physics Leters, 88 (2006), 241910-1-241910-3. 12. T. J. Richardson, B. Farangis, J. L. Slack, P. Nachimuthu, R. Perera , N. Tamura, and M. Rubin, "X-ray absorption spectroscopy of transition metal magnesium hydride thin films," Journal ofAlloys and Compounds, 356-357 (2003), 204-207. 13. P. Vermeulen, R. A. H. Niessen, and P. H. L. Notten, "Hydrogen storage in Metastable MgyTi(i.y) thin films," Electrochemistry Communications, 8 (2006), 27-32.
Acknowledgements The authors thank M. Tessema for assistance with XRD measurements and R. Waldo for EPMA measurements.
14. P. Vermeulen, R. A. H. Niessen, D. M. Borsa, B. Dam, R. Griessen, and P. H. L. Notten, "Effect of the deposition technique on the metallurgy and hydrogen storage characteristics of metastable MgyTi(1.y) thin films," Electrochemical and Solid-State Letters, 9 (2006), A520-A523.
References 1. K. R. Baldwin, D. J. Bray, G. D. Howard, and R. W. Gardiner, "Corrosion behaviour of some vapour deposited magnesium alloys," Materials Science and Technology, 12 (1996), 929-944.
15. G. Song, and D. Haddad, "The topography of magnetron sputter-deposited Mg-Ti alloy thin films," Materials Chemistry and Physics, accepted to be published.
2. Y. Bohne, D. M. Seeger, C. Blawert, W. Dietzel, S. Mändl, and B. Rauschenbach, "Influence of ion energy on properties of Mg alloy thin films formed by ion beam sputter deposition," Surface & Coatings Technology, 200 (2006), 6527-6532. 3. T. Mitchell, S. Diplas, and P. Tsakiropoulos, "Characterisation of corrosion products formed on PVD in situ mechanically worked Mg-Ti alloys," Journal of Alloys and Compounds, 392 (2005), 127-141.
16. W. C. Oliver and G. M. Pharr, "An improved technique for determining hardness and elastic modulus using load and displacement sensing indentation experiments," Journal of Materials Research, 7 (1992), 1564-1583. 17. R. A. Waldo, M. C. Militello, and S. W. Gaarenstroom, "Quantitative thin-film analysis with an energy-dispersive x-ray detector," Surface and Interface Analysis, 20 (1993), 111-114.
4. S. Rousselot, M. P. Bichat, D. Guay, and L. Roue, "Structure and electrochemical behaviour of metastable Mg50Ti5o alloy prepared by ball milling," Journal of Power Sources, 175 (2008), 621-624.
18. L. Vegard, "Die Konstitution der Mischkristalle und die raumfüllung der atome," Zeitschrift für Physik, 5 (1921), 17-26.
5. Y. T. Cheng, M. W. Verbrugge, M. P. Balogh, D. E. Rodak, and M. Lukitsch, "Magnesium-Titanium solid solution alloys," United States Patent, US 7651732B2, Jan. 26, 2010.
19. M. S. Bobji, S. K. Biswas, and J. B. Pethica, "Effect of roughness on the measurement of nanohardness -A computer simulation study," Applied Physics Letters, 71 (1997), 1059-1061.
6. J. L. Murray, "The Mg-Ti (Magnesium-Titanium) System," Bulletin ofAlloy Phase Diagrams, 7(1986), 245-248.
20. J.Y. Kim, J.J. Lee, Y.H. Lee, J.I. Jang, and D. Kwon, "Surface roughness effect in instrumented indentation: A simple contact depth model and its verification," Journal of Materials Research, 21 (2006), 2975-2978.
7. W. P. Kalisvaart, H. J. Wondergem, F. Bakker, and P. H. L. Notten, "Mg-Ti based materials for electrochemical hydrogen storage," Journal ofMaterials Research, 22 (2007), 1640-1649. 8. R. A. H. Niessen and P. H. L. Notten, "Electrochemical Hydrogen Storage Characteristics of thin film MgX (X = Sc, Ti, V, Cr) Compounds," Electrochemical and Solid-State Letters, 8 (2005), A534-A538.
621
Magnesium Technology 2011 Edited by: Wim H. Sillekens, Sean R. Agnew, Neale R. Neelameggham, andSuveen N. Mathaudhu TMS (The Minerals, Metals & Materials Society), 2011
Effect of Adding S1O2-AI2O3 Soi into Anodizing Bath on Corrosion Resistance of Oxidation Film on Magnesium Alloy Huicong Liu, Liqun Zhu, Weiping Li Key Laboratory of Aerospace Materials and Performance (Ministry of Education), School of Materials Science and Technology, Beihang University, 37# Xueyuan Road, Haidian District, Beijing, 100191, China Keywords: magnesium alloy, modified oxidation film, Si02-Al203 sol, anodic oxidation alloy under the effect of silica sol[8,9,17] were investigated, and the effect of sol composition on anodic oxidation film on magnesium alloys[21] was studied. The results showed that the addition of silica sol in Na2Si03 solution could decrease the surface energy and the conductivity of the solution, and could increase the anodic film thickness and improve the uniformity of the anodic film on AZ91D magnesium alloy.
Abstract Due to the widely use in automobile and construction field, AZ91D magnesium alloy need to be protected more effectively for its high chemical activity. In this paper, three kinds of films were formed on magnesium alloy. The first kind of film, named as anodic oxidation film, was prepared by anodic oxidation in the alkaline solution. The processes for preparing the second kind of film, named as multiple film, involved coating sol-gel on the samples and heat-treating before anodic oxidation. The third kind of film was prepared by anodic oxidation in the alkaline oxidation solution containning 5% (vol) Si02-Al203 sol, named as modified oxidation film. The corrosion resistance of the three different films was investigated. The results showed that the modified oxidation film had the highest corrosion resistance due to the largest thickness and most dense surface morphology. Sol was discussed to react during the film forming process, which leaded to the difference between modified oxidation film and anodic oxidation film.
In this paper, modified oxidation film was prepared by adding sol into alkaline oxidation solution, and its corrosion resistance was investigated in comparison with multiple film and anodic oxidation film. The purpose is to form a more protective film on magnesium alloy surface using more simple process. Experimental Procedure (1) Materials The raw material was AZ91D molten magnesium alloy, and samples were cut into the size of 40mmx25mmx2mm. The chemical composition(mass percentage) includes: 8.5 ~ 9.5 % Al,
Introduction Magnesium alloy has been widely used due to its advantages such as low density, high strength-to-density ratio, nice electromagnetic shielding, etc[l-2]. However, magnesium alloy is very active and its oxidation film formed in air is very loose and porous[3], therefore, magnesium alloy is very apt to be corroded in corrosive environment.
0.17-0.4% Mn, 0.45 - 1.90% Zn, 0.05 % Si, 0.25% Cu, 0.001 % Ni, 0.004% Fe and Mg is the balance. (2) Preparation of films on magnesium alloy Before the films were prepared, the magnesium alloy samples were polished using sand paper and then degreased in acetone in an ultrasonic device.
In order to protect magnesium alloy, many methods have been used, such as chemical conversion[4-6], anodic oxidation[7-ll], micro-arc oxidation[12-13], electroless plating[14], plating[ 15-16] and so on. Among these methods, anodic oxidation has been studied by many researchers and has made some progress. However, the forming oxidation film on magnesium alloy has many pores[17-18], which will lead to the film easy to be corroded after some certain time.
The prepared sol was Si02-Al203 composite product (volume rate is 70:30). The reactant used to prepare sol was methyltriethoxysilane and aluminum iso-propoxide. The solvent was absolute ethanol and acetone. The reaction time was 6~8 hours, and the temperature was 60~90°C. The power supply for anodic oxidation process was an autotransformer (voltage and current output are 0~250V and 0-4.0A, respectively).
In recent years, sol-gel dip coating method has been used to increase corrosion resistance of magnesium alloy, aluminum alloy and nickel plating on carbon steel[ 18-20], and sol plays a role in sealing holes on surface coating.
Three kinds of films were made in this paper. The process for preparing the first kind of film was anodic oxidation in the solution containing 150g/L NaOH, 35 g/L Na3P04 and 35 g/L NaF. The anodic oxidation voltage was 65V and the duration was 30 min. This kind of film is named as "anodic oxidation film".
Generally speaking, sol-gel film is very thin and several dip coating cycles are needed to obain expected thickness. In order to get excellent film through more simple process, sol was added into anodizing bath and was expected to plays a positive role in anodizing process.
The second kind of film was made by coating sol on the samples and heat-treating before anodic oxidation. The parameters for
In our previous study, electrochemical oxidation characteristic and growth characterization of anodic film on AZ91D magnesium
623
anodic oxidation process were the same to that of anodic oxidation film. This kind of film is named as "multiple film".
kinds of films. It can be seen that compared with anodic oxidation film and multiple film, modified oxidation film has the lest corrosion pit number after dipped in the 5wt% NaCl solution for 12h in the same area, and can stand the longest time before being corroded. In other word, it owns the highest corrosion resistance.
The third kind of film was prepared by anodic oxidation, the oxidation solution was prepared by adding 5%(vol) Si02-Al203 sol into the anodic oxidation solution. The anodic oxidation voltage and duration were the same to that of anodic oxidation film. And this kind of film is named as "modified oxidation film".
Tablel
Corrosion Resistance of the Three Kinds of Films
Films
(3) The device and methods for measurement Corrosion pit number (/10cm2) Time for pit emerging (h)
The thickness of the films was measured by E110B whirlpool testing instrument made by FISCHER company in west Germany. The surface morphology was analyzed by S-530 scanning electron microscope and Hirox KH-3000 digital microscope. The composition of films was examined by Link ISIS instrument made in Japan.
Anodic oxidation film 8-10
Multiple film 9-11
Modified oxidation film 1-2
4-6
3-5
12-14
Figurel I is the result of galvanic corrosion test for three kinds of films. It can be seen that in the same corrosive condition, modified oxidation film has the lowest current value in comparision with the other two kinds of films. From figure 111, there are fewer corrosion pits on the surface of modified oxidation film than that of anodic oxidation film, which further comfirms the results in figurel and II .
The corrosion resistance was studied in 5 wt% sodium chloride solution, and the techniques used included salt-solution dipping, anodic polarization curve measurement and galvanic corrosion test. The anodic polarization curves were recorded on a CHI604A electrochemical analyzer. In salt-solution dipping test, the anticorrosion time was defined as the time of the first corrosion spot to be found. The galvanic corrosion test was done on ZRA-1 galvanic corrosion instrument. Galvanic corrosion current was examined relative to LY12CZ aluminum alloy.
From the three kinds of corrosion test, the same conclusion can be achieved that the modified oxidation film has the highest corrosion resistance.
Results and discussion
(2) Discussion
(I) Corrosion resistance of modified oxidation film
As shown in tablell , the modified oxidation film has the largest thickness in comparision with anodic oxidation film and multiple film. And the thickness of multiple film is a little smaller than that of anodic oxidation film. Figure IV shows the cross section morphology of modified oxidation film, which matches well with the thickness of this film in tablell .
As shown in figure I , the modified oxidation film has much higher corrosion resistance than that of anodic oxidation film and multiple film. Figurel shows that the potential of the samples coated with modified oxidation film is more positive than those of anodic oxidation film and multiple film at the same current value, and the largest potential difference for the examined value is about 850mV, which indicates that the samples treated with modified oxidation film have small tendency to be corroded than the other two kinds of films in the same condition.
70 _
60 50
J< / '
< 40 E S 30 20 10
f»
1f
i/
*
m
^^r^V-->% • ^*>~ ••' f
1
7
* —□— multiple film - v - modified oxidation film • o« anodic oxidation film
10 time / min
15
20
Figll Corrosion current trends for three kinds of films Figl
Anodic polarization curves of three kinds of films
The result of salt-solution dipping test is presented in tablel , which shows the corrosion resistance comparison of the three
624
film, as shown in tablel I . Due to the breaking and solving effect, the anodic oxidation process would be affected, which leads to a slightly thinner film. After adding 5% (vol) Si02-Al 2 0 3 sol into anodic oxidation solution, the thickness of oxidation film increased greatly and is much larger than those of anodic oxidation film or multiple film. This may be due to that the sol in the anodic oxidation solution greatly affected the oxidation process, and increased the growing rate of oxidation film. This can be verified in tablelll. It can be seen that Si and Al elements were found in the modified oxidation film, while the Al element content is lower than that of anodic oxidation film or multiple film. The reason is not very clear yet, and will be studied in detail in future work. Tablel 11 Element Composition Comparison of the Films (Atom%) Si Films O Al Mg anodic oxidation 27.1-29.2 65.5-68.5 4.1-5.7 film
Figlll Micro morphologies of magnesium alloys samples after galvanic corrosion test with aluminum alloy for 15min Tablel I Film Thickness(u,m)
Thickness of the Three Kinds of Films Anodic oxidation film 4.0-4.9
Multiple film 3.4-4.0
Modified oxidation film 29.8-36.2
multiple film modified
26.8-27.7
65.9-68.3
5.0-6.3
oxidation film
29.8-33.6
63.5-64.9
2.3-3.2
1.8-2.4
In the process of anodic oxidation in the solution containing sol, sol transfers to magnesium alloy surface. Thanks to the exothermic action of sparks discharging, the sol network is destroyed. The electriferous groups such as -S\- , :AI • , : Si - R ( alkyl ) ; :AI - R ( alkyl ) , : Si - 0 > : A | _ 0 w i u form and adsorb on the surface, which participate in the oxidation reaction and forming film, and hence Si0 2 and A1 2 0 3 formed. These reactions will increase the growth rate of oxidation film on magnesium alloy, so there is an obvious increase in film thickness. So the film can act as barrier effect and increase corrosion resistance of modified oxidation film. Although the comparision experiments have not been done with the addition of inorganic salt, the effect of sol on forming modified oxidation film is obvious. Comparing the morphology of modified oxidation film(figureV c) with those of anodic oxidation film(figureV a) and multiple
FiglV Cross section morphology of modified oxidation film
film(figureV b), it can be seen that there are some relatively big holes and some holes connect with each other on the anodic oxidation film and multiple film. These holes will obviously reduce the corrosion resistance of magnesium alloy. In comparision with the anodic oxidation film and multiple film, the modified oxidation film is uniform and dense and no big holes on the film are observed, which can effectively increase the corrosion resistance of oxidation film.
The multiple film is thinner, perhaps because there is a breaking and solving effect on the formed sol-gel film on the surface during the anodic oxidation process. Because the sol film is rather thin, it can be equally broken down when the applied voltage is high enough for the following anodic oxidation process, then sol would go into the oxidation solution. As a result, the composition of the multiple film has little difference with that of anodic oxidation
With respect to anodic oxidation process, a common viewpoint is that at the initial stage of oxidation, when voltage is higher than the broken voltage of the oxidation film formed in air, sparkle discharging phenomenon would appear. Since the instantaneous temperature induced from sparkle discharging is very high, magnesium and other alloy composition would be partially melted and then oxidation film would be formed under the cooling effect of solution. As oxidation duration prolongs, voltage to break the
sample surface film persistently increased, and the repeating break of formed film would lead to the increase in film thickness, and thus oxidation process could go on. A great deal of heat from sparkle discharging was absorbed by solution, and metal oxide formed on sample surface was cooled, which leads to the shrinkage of formed film and hence porous morphology of anodic oxidation film was obtained.
figurelV, there are no penetrable holes in the film, this denser and thicker film would increase the corrosion resistance of magnesium alloy.
Conclusion (1) After adding Si02-Al203 sol into anodizing bath, modified oxidation film has the largest thickness of about 34um and the highest corrosion resistance relative to anodic oxidation film and multiple film. In the 5wt% NaCl solution, time for pits emerging of modified oxidation film was about three times longer than that of anodic oxidation film. (2) The modified oxidation film had the largest thickness, was the most dense and had the least pinholes on the surface among the three kinds of films, and no penetrable holes were observed from the cross section morphology. These were the reasons for the highest corrosion resistance. (3) Sol played an important role in film forming process, some electriferous groups were formed under the condition of high temperature and electric intensity. These groups adsorbed on the sample surface and reacted to form film. This process leaded to a more thick and dense modified oxidation film in comparision with anodic oxidation film. Acknowledgement This work is supported by the Fundamental Research Funds for th e Central Universities(YWF-10-02-036), the Cheung Kong Schola rs and Innovative Research Team Program in University from Mi nistry of Education ( Grant No. IRT0805 ) and the aerial science fund (2008ZE51064). References [1] Yan A.J., Zhu X.M., Teng Y., etc. "Electrochemical Corrosion Behavior of MAO Film on AZ31 Magnesium Alloy"[J], Journal of Dalian Jiaotong University, 2008, 29(3):45-48. [2] R. Arrabal, E. Matykina, T. Hashimoto, etc. "Characterization of AC PEO Coatings on Magnesium Alloys"[J], Surface and Coatings Technology, 2009, 203(16):2207-2220. [3] Liu Y.G., Zhang W., Li J.Q.. "Microarc Electrodeposition of Ceramic Coatings on Double Electrodes of AZ91D Magnesium Alloy by AC Pulse Method"[J], Journal of university of science and technology Beijing, 2004 , 26(1) : 73-77. [4] J.K. Lin, J.Y. Uan. "Formation of Mg,Al-hydrotalcite Conversion Coating on Mg Alloy in Aqueous HC037C032~ and Corresponding Protection Against Corrosion by the Coating"[J], Corrosion Science, 2009, 51(5): 1181-1188. [5] Huo H.W., Li Y., Wang F.H.. "Effect of Chemical Conversion Film Plus Electroless Nickel Plating on Corrosion Resistance of Magnesium Alloys"[J], The Chinese Journal of Nonferrous Metals, 2004, 14(2):267-272. [6] Zhao M., Wu S.S., Luo J.R., etc. "The Present Status and Prospect of Chromium Free Surface Treatment for Magnesium Alloys"[J], Foundry, 2003, 52(7):462-465. [7] A. Yabuki, M. Sakai. "Anodic Films Formed on Magnesium in Organic, Silicate-containing Electrolytes"[J], Corrosion Science, 2009, 51(4): 793-798.
FigV Micro morphologies of the three kinds of films The effect of sol on the thickness and density of the oxidation film may be explained as follows. The sol in anodic oxidation solution will decrease the conductivity of solution, as reported in reference 17. At the early stage of anodic oxidation, the spark discharging voltage increased, higher than that in anodic oxidation solution, then the growth rate of oxidation film increased and the rate of holes in the film also increased. When anodic oxidation went on, Si02-Al203 sol in the anodizing solution reacted on the interface of anodic oxidation film, and some Si02-Al203 sol composition entered into the film. So the film was more dense and there were fewer big holes. With oxidation duration increasing, the anodic oxidation process continued, and due to sol composition act on the interface of the film or sealed some defaults, the film growed gradually and the thickness of the film increased. The more dense and uniform modified oxidation film will get the better corrosion resistance than anodic oxidation film and multiple film. And from
626
[8] Li W.P., Zhu L.Q., Li Y.H.. "Electrochemical Oxidation Characteristic of AZ91D Magnesium Alloy under the Action of Silica Sol"[J], Surface and Coatings Technology, 2006, 201(34): 1085-1092. [9] W.P. Li, L.Q. Zhu, Y.H. Li, etc. "Growth Characterization of Anodic Film on AZ91D Magnesium Alloy in an Electrolyte of Na2SiÛ3 and KF'[J], Journal of University of Science and Technology Beijing, Mineral, Metallurgy, Material, 2006, 13(5): 450-455. [10] H. Asoh, S. Ono. "Anodizing of Magnesium in Amineethylene Glycol Electrolyte"[J], Materials Science Forum, 2003(419-422):957-962. [11] L.Y. Chai, X. Yu, Z.H. Yang, etc. "Anodizing of Magnesium Alloy AZ31 in Alkaline Solutions with Silicate under Continuous Sparking"[J], Corrosion Science, 2008, 50(12): 3274-3279. [12] P. Shi, W.F. Ng, M.H. Wong, etc. "Improvement of Corrosion Resistance of Pure Magnesium in Hanks' Solution by Microarc Oxidation with Sol-Gel Ti0 2 Sealing"[J], Journal of Alloys and Compounds, 2009, 469(1-2): 286-292. [13] Liu Y.G., Zhang W., Li J.Q.. "Microarc Electrodeposition of Ceramic Films on Double Electrodes of AZ91D Magnesium Alloy by Symmetrical AC Pulse Method"[J], Surface Engineering, 2003, 19(5):345-350. [14] Huo H.W., Li Y., Wang F.H.. "Electroless Nickel Plating on AZ91D Magnesium Alloys"[J], Journal of Chinese Society for Corrosion and Protection, 2002, 22(1): 14-17. [15] Cao W.B., Ren C.X., Guan S.K., etc. "Research Progress in the Electric Ni-P on Magnesium Alloys"[J], Water Conservancy and Electric Power Machinery, 2003, 25(4):28-31. [16] J.F. Zhang, C.W. Yan, F.H. Wang. "Electrodeposition of AlMn Alloy on AZ31B Magnesium Alloy in Molten Salts"[J], Applied Surface Science, 2009, 255(9):4926-4932. [17] W.P. Li, L.Q. Zhu, H.C. Liu. "Effects of Silicate Concentration on Anodic Films Formed on AZ91D Magnesium Alloy in Solution Containing Silica Sol"[J], Surface and Coatings Technology, 2006, 201(6): 2505-2511. [18] Cai Q.Z., Wang D., Luo H.H., etc. Sealing of Micro-arc Oxidation Coating on Magnesium Alloy by Si0 2 Sol Sealing Agent[J], Special Casting & Nonferrous Alloys, 2006, 26(10):612615. [19] Zhou Q., He CL., Cai Q.K., etc. Sealing of Anodized Films on Al Alloy with Boehmite Sol[J], The Chinese Journal of Nonferrous Metals, 2007, 17(8): 1386-1390. [20] Liu H.C, Zhu L.q., Du Y.B.. High Temperature Oxidation Resistance of Thin Film Made by Sol-gel Method[J], Trans Mater Heat Treat, 2004, 25(4):77~80. [21] Zhu L.Q., Liu H.C. "The Effect of Sol Ingredient to Anodic Oxidation Film on Magnesium Alloys"[J], Journal of functional materials, 2005, 36(6):923-926.
Magnesium Technology 2011 Edited by: Wim H. Sillekens, Sean R. Agnew, Neale R. Neelameggham, andSuveen N. Mathaudhu TMS (The Minerals, Metals & Materials Society), 2011
Monotonie and Fatigue Behavior of Mg Alloy Friction Stir Spot Welds: An International Benchmark Test in the "Magnesium Front End Research and Development" Project H. Badarinarayan1, SB. Behravesh2, S.D. Bhole3, D.L. Chen3, J. Grantham4, M.F. Horstemeyer4, H. Jahed2, J.B. Jordon5, S. Lambert2, H.A. Patel3, X. Su6, and Y. Yang7 2
1 Automotive Products Research Laboratory, Hitachi America Limited, USA Mechanical and Mechatronics Engineering Department, University of Waterloo, Canada department of Mechanical and Industrial Engineering, Ryerson University, Canada 4 Center for Advanced Vehicular Systems (CAVS), Mississippi State University, USA 'Department of Mechanical Engineering, The University of Alabama, USA 6 Ford Motor Company, USA 'institute for Metals Research, China
Keywords: Fatigue, Friction stir spot welds; Magnesium alloys; Resistance spot weld targeted as a possible joining technique in the fabrication of a front end of an automobile using magnesium alloys.
Abstract This paper presents the experimental results of benchmark coupon testing of monotonie and cyclic conditions on friction stir spot welded coupons of Mg AZ31 alloy. The results presented here are a product of a collaborative multinational research effort involving research teams from Canada, China, and the United States. Fatigue tests were conducted in load control at R=0.1 at two different maximum loads: lkN and 3kN. Good agreement was found between the participating labs regarding the number of cycles to failure. Differences in the failure modes were observed for the two different loading conditions tested. At the higher load, fatigue failure was caused by interfacial fracture. However, at the lower load, fatigue cracks formed perpendicular to the loading direction, which led to full width separation. For additional comparison, the monotonie and cyclic results of the friction stir spot welds are compared to resistance spot welded coupons of similar nugget size.
Friction stir welding has steadily been gaining more wide spread use when high integrity and strength are required. Friction stir spot weld is a recent variant of friction stir welding and is an attractive welding technique due to the solid state nature of the process and the lack of stress relieving that is typically needed. A recent literature review of the friction stir spotfrictionprocess can be found in [8]. The solid state nature and the ability of joining dissimilar metals have madefrictionstir spot welding an attractive welding process. The fatigue behavior of a friction stir spot weld (FSSW) is highly dependent on process parameters employed to create the weld [9,10]. These parameters include speed, depth of plunge, dwell time, and tool configuration. Up to this point, a majority of studies of the nature of fatigue of FSSW's have been almost exclusively focused on aluminum alloys [9-12], With regards to FSSW's made of aluminum alloys, the failure modes of quasi-static and cyclic loads of FSSW coupons vary based on the load level [9], Differences have been observed in the fracture path for quasistatic loads compared to fatigue loads for aluminum alloys [9]. In this study[9-10], FSSW coupons failed by interfacial fracture under quasi-static loading, whereas, under cyclic loading, fatigue cracks initiated and grew from several locations including the interfacial tip and outside the weld zone. In addition, fatigue failure modes of aluminum FSSW made using various tooling were also observed to differ based on the shape of the tooling used to make the weld [10]. Also, fatigue of FSSW's of dissimilar aluminum alloys were observed to have different failure modes compared to FSSW's with identical alloys for the top and bottom sheets [11]. The fatigue failure modes of dissimilar metals (aluminum and steel) were also observed to vary compared with joints made of all aluminum alloys [12]. While the FSSW joining technique of magnesium alloys is documented [13-16], limited published literature exists on the fatigue properties of FSSW coupons of magnesium alloys. Mallick and Agarwal [17] were the first to quantify the fatigue behavior of FSSW's made of a magnesium alloy. However, their characterization of the failure mechanisms under cyclic loading was limited in its presentation.
Introduction The continued push for more fuel efficient automobiles designs is motivation for ongoing research in lightweight metals. Through lightweight designs comes the need to explore alternatives to traditional metals like aluminum and steel currently used in the manufacturing of automobiles. As such, a collaborative multinational research effort involving researchers from Canada, China, and the United States and joined by Chrysler, Ford, and General Motors, is underway with the goal of developing the ability to build a front end of an automobile constructed of magnesium alloys. Magnesium alloys, with a density of 1.74g/cm3, weigh less than a quarter of steel and two-third of aluminum, and with abundant reserves on the earth, are attractive substitutes for steel and aluminum for vehicle body structures [14]. Replacing the largely steel structure of a vehicle's front end with magnesium can also move the vehicle center of gravity away from the front, improving vehicle drivability. This research project, called the Magnesium Front End Research and Development program, or MFERD, has already yielded several fatigue characterizations of wrought magnesium alloys [cf. 5-7], In order to fully meet the project's objectives, characterization of mechanical properties of potential joining techniques is also needed. As such, friction stir spot welding is
Resistance Spot Weld (RSW), on the other hand, is currently the most common joining process in the automotive industry and
629
hence is the manufacturers' preferred joining technique. Thus, as have been performed for steel [18-22] and aluminum alloys [2327], characterization of FSSW and RSW techniques and comparing their monotonie and cyclic behavior, provide the motivation for these types of studies. As such, the purpose of this research is the characterization of FSSW'ed sheets of AZ31B magnesium alloy through a round-robin experimental program and also evaluating its mechanical behavior by a comparative analysis to RSW of the same alloy.
two (2) Pmax load levels: 3kN and lkN. The tests were run alR = _E!2. = o.l, and a frequency of 5 Hz. All tests were •max
conducted at ambient temperature and humidity. The test set-up also included certain specifications for mounting including careful attention to grip alignment, grip distance, and use of shims or offsets. The grip-to-grip distance was maintained at 110 mm at each of the labs. In addition, UW also tested RSW coupons under the same conditions as stated for the FSSW coupons.
Materials
Monotonie Tests
Magnesium AZ31B alloy sheets of 2.0mm thickness are chosen for the present study. Coupons were welded in lap configuration. The individual sheet dimensions were: length 100mm, width 38mm and were welded on an overlap area of 38 x 38 mm. The FSSW tool was made from standard tool steel (H13) material, having a shoulder with diameter 12 mm, pin length of 3.2 mm and left hand threads (M5). The shoulder was a concave profile with the angle of concavity of 10 deg. The welding process parameters were: tool rotation speed 750 RPM, tool plunge speed of 20mm/min, shoulder plunge depth of 0.1mm and a dwell time of 2.5 sec. In regards to the bonded area in FSSW, the welding process produces an annulus shape, and the average inner and outer diameters in this study were found to be 6.5mm and 9.7mm, respectively. During the welding process, the interface between the upper and lower sheet is formed into a hook like shape due to the penetration of the FSSW tool into the bottom sheet.
Results and Discussion
Monotonie tests were performed on a servo-hydraulic load frame under uniaxial displacement rate of 1 mm/min. The test results show that the FSSW coupons exhibited an average ultimate tensile-shear load (UTSL) of 4650±10N. All specimens tested failed in a consistent manner with a partially interfacial failure mode, as shown in Figure 2.
Figure 2. Failure mode in friction stir spot welded coupons under monotonie tensile-shear loading. Similar to studies on steel FSSW [22] and aluminum FSSW [27], comparing the static and cyclic behavior of emerging FSSW joining technique and commonly used RSW is of interest. Recent research [28] has studied the monotonie and fatigue behavior of RSWs of AZ31B-H24 Mg alloy, and showed that the highest UTSL is achieved using the welding current of 34 kA, and the welding time of 8 cycles. The same welding parameters were utilized in this study, and a solid circular nugget with an average diameter of 10.4 mm was obtained which is close to outer diameter of FSSW (9.7 mm). It should be noted that although FSSW and RSW are of a similar outer diameter, the bonded area are very different (40 mm2 in FSSW and 85 mm2 in RSW), due to the different shapes of bonded region. However, comparing the mechanical behavior of these specimens is sound from the application perspective, as the area of coupons contributing in the joints is almost the same in FSSW and RSW specimens. Monotonie testing of tensile-shear RSW specimens yielded an average UTSL of 7620+48N with interfacial failure mode, as shown in Figure 3.
For comparison purposes, resistant spot weld (RSW) coupons were made of the same alloy as for the FSSW. The thickness of the sheet as well as the specimen configuration and dimensions were similar to the FSSW. Additionally, the weld nugget was approximately the same. The RSW parameters were: welding current of 34 kA, welding time of 8 cycles (8/60 sec), electrode force of 4 kN, and holding time of 30 cycles (0.5 sec). Figure 1 shows general dimensions for FSSW and RSW coupons employed in this study.
Figure 3. Failure mode in resistant spot welded coupons under monotonie tensile-shear loading.
Figure 1. Configuration of friction stir and resistant spot welds single-weld lap-shear coupons. Dimensions are in mm.
Load-displacement curves, shown in 4, illustrate that the RSW coupons have higher UTSL compared to FSSW coupons. Higher UTSL of RSW specimens is mainly attributed to the smaller bonded area in FSSW specimens. As mentioned before, RSWs have a solid circular, and FSSWs have an annulus-shaped weld region. Therefore, even for the same outer diameter, a higher UTSL is expected for RSWs.
Experiments Mississippi State University (MSU), Ryerson University (RU), University of Waterloo (UW), and the Institute of Metal Research (IMR), all participated in round-robin fatigue testing of lap-shear FSSW coupons. Each institution conducted six (6) load control tests consisting of three (3) specimens conducted at each of the
630
to Kr in coupon separation is parallel to the coupon surface or normal to the coupon thickness at the joint. And because the load component parallel to the coupon is larger than the one normal to the coupon surface (as bending rotation in coupon at the joint is less than 45 deg), the coupon separation mode is more critical than interfacial failure. However, this still needs more investigations.
Figure 4. Load-displacement curves for friction stir spot weld and resistance spot weld tensile-shear specimens under quasi-static loading. Fatigue Tests Figure 5 displays the round-robin fatigue results of the FSSW from the four institutions. The first observation based on the fatigue life results was the overall good correlation between all of the laboratories. Fatigue failure for this particular study was defined as complete separation of the lap-joint. However, there were some outliers in both sets of load levels tested. One of the tests MSU conducted at the 3kN load level failed much earlier than the rest of the tests compared to the other institutions. At the lkN level, several specimens tested at IMR failed much later compared to the rest of the group of coupons. However, the variation in the fatigue results overall is fairly consistent compared to other published round-robin fatigue testing programs. It is important to point out that at the higher fatigue load level (3kN), the load is above the elastic limit based on the monotonie load-displacement curve shown in Figure 4. At the lkN load level, the load is within the elastic range. The number of cycles to failure for the FSSW coupons presented here are in the same order of magnitude of similar FSSW [9-12,17]
Figure 5. Comparison of Mg AZ31 friction stir spot welds fatigue results from the four universities: Mississippi State University (MSU), Ryerson University (RU), University of Waterloo (UW), and Institute for Metal Research (IMR). Fatigue tests were conducted in load control at R=0.1, at a frequency of 5 Hz and at room temperature.
For further comparisons, the fatigue life of the FSSW coupons are compared to the fatigue life of the RSW. Fatigue results show that the RSW exhibited better fatigue life at the lower load level (lkN) compared to the FSSW coupons. However at the higher load level (3kN), the FSSW coupons exhibited better fatigue resistance compared to the RSW coupons. For the lower load level (lkN), where we have the same failure modes in FSSW and RSW specimens (coupon width separation, Fig. 7.a), higher fatigue life of RSW could be attributed to larger nugget size which causes lower stress concentration and hence smaller hot spot stress and retardation of crack initiation. However, at the higher load level (3kN), the better fatigue performance of FSSW is due to the different failure modes in FSSW and RSW specimens. FSSW specimens failed in interfacial mode (Fig. 7.b), while coupon failure perpendicular to the loading direction (Fig. 7.a) was observed in RSW specimens. The reason why the fatigue strength in interfacial mode is higher than coupon failure is, in both cases, the mode I stress intensity factor (Kr) is the main factor for fatigue crack propagation. The load component contributing to Ki in interfacial failure mode is normal to the coupon surface at the joint, and the load component contributing
Figure 6. Comparison of fatigue results of the magnesium AZ31 alloy friction stir spot welds to the resistance spot welds tested at University of Waterloo (UW). Fatigue tests were conducted in load control at R=0.1, at a frequency of 5 Hz and at room temperature.
631
Fractographv The failure modes under cyclic loading were observed to vary for the different load levels tested in the round-robin testing program. All four laboratories reported that the failure mode at the lower cyclic load level of 1 kN was different compared to the higher cyclic load level of 3 kN. That is, the failure at the lower load level occurred perpendicular to the loading direction, while the failure at the higher load level exhibited interfacial fracture, as shown in Figure 7(a) and (b), respectively. The interfacial failure observed for the 3 kN load level likely occurred because as the crack propagated circumferentially around the nugget, the shear/tensile stress in the remaining net area of the nugget increased with each advancement of the crack front. Once the crack had propagated around approximately half of the nugget diameter, the shear/tensile stresses acting on the net area were such that the remaining cross section failed under shear/tensile overload. The other type of fatigue failure occurred at the load level of lkN. Once the crack had propagated circumferentially around the nugget, the crack then propagated outward through the sheet material.
Figure 9. Scanning electron microscope images of fracture surfaces of friction stir spot weld coupons of magnesium AZ31 alloy fatigued at a load level of Pmax=3 kN, (a) overall view of interfacial fracture of the sample, (b) initiation site in the boxed region in (a), (c) magnified view near the initiation site, and (d) crack propagation area at a higher magnification. Fracture surfaces of the fatigued specimens were examined under scanning electron microscope (SEM). Figure 8(a-d) and Figure 9(a-d) show the typical SEM images of the coupons tested at Pmax=l kN and Pmax=3 kN, respectively. The low magnification image shown in Figure 8(a) was taken near the center of the sample fractured at Pmax=l kN, while 9(a) showed an overall view of interfacial fracture of the sample tested at Pmax=3 kN. Figure 8(b) and Figure 9(b) showed the boxed region in Figure 8(a) and 9(a), respectively, indicating the crack initiation site. While the failure mode was different (see Figure 7), the fatigue crack initiation at both load levels occurred from the surface. Figure 8(c) and 9(c) showed the higher magnification images near the fatigue crack initiation sites for both load levels. Figure 9(d) shows the fatigue crack propagation area with some striation-like features perpendicular to the crack propagation direction.
Figure 7. Macroscopic images of friction stir spot weld coupons of magnesium AZ31 alloy samples fatigued at a load level of (a) Pmax=l kN, and (b) Pmax=3 kN (R=0.1, 5 Hz, sine waveform, room temperature).
Conclusions A summary of the main conclusions of this work are as follows:
Figure 8. Scanning electron microscope images of fracture surfaces of friction stir spot weld coupons of magnesium AZ31 alloy fatigued at a load level of Pmax=l kN, (a) low magnification image near the center of the sample, (b) initiation site in the boxed region in (a), (c) magnified view of the boxed region in (b), and (d) crack propagation area near the center of the sample thickness.
632
1.
The resistant spot weld lap-shear coupons exhibited better monotonie strength compared to the friction stir spot weld coupons with similar specimen dimensions and outer nugget diameter.
2.
Friction stir spot weld and resistant spot weld coupons both failed in the interfacial mode under monotonie loading.
3.
The fatigue results from the four different testing labs demonstrated consistent fatigue results on the friction stir spot weld lap-shear coupons.
4.
Different failure modes were observed for the friction stir spot weld coupons for the two cyclic load levels tested. At the high cyclic load level (3kN), the coupons failed by interfacial fracture. At the lower cyclic load level (lkN), the coupons failed by full width separation.
5.
Striations-like features were observed on fracture surfaces of specimens tested at 3kN.
6.
The fatigue life the friction stir spot weld coupons were observed to compare closely to the fatigue life of the resistant spot weld coupons. At the higher cyclic load (3kN), the friction stir spot weld coupons exhibited better fatigue life, compared to the resistant spot weld coupons. However, at the lower cyclic load (lkN), the resistant spot weld coupons exhibited better fatigue resistance compared the friction stir spot weld coupons.
of Aluminum 6111-T4 Sheets. Part 2: Welds Made By a Flat Tool," Int. J. Fatigue, 30(1), pp. 90-105. 11. Tran, V.-X., Pan, J., Pan, T., 2010, "Fatigue behavior of spot friction welds in lap-shear and cross-tension specimens of dissimilar aluminum sheets," Int. J. Fatigue, 32(7), pp. 1022-1041 12. Tran, V.-X., Pan, J., 2010, "Fatigue behavior of dissimilar spot friction welds in lap-shear and cross-tension specimens of aluminum and steel sheets," Int. J. Fatigue, 32(7), pp. 1167-1179 13. Su, P., Gerlich A., and North, T.H., 2005,"Friction stir spot welding of aluminum and magnesium alloy sheets, Society of Automotive Engineers," Warrendale (PA) (2005) [SAE Technical Paper No. 2005-01-1255. 14. Gerlich, A., Su, P., and North, T.H., 2005, "Tool penetration during friction stir spot welding of Al and Mg alloys," J Mater Sei 40 pp. 6473-6481. 15. Pan, T.-Y., Santella, M., Mallick, P.K., Frederick, A., Schwartz, W.J., 2006, "A feasibility study on spot friction welding of magnesium alloy AZ31," In: Proceedings of 63rd annual world magnesium conference, Beijing, China, May 21-24; pp. 179-86. 16. Agarwal, L., Mallick, P.K., and Kang, H.T., 2008, "Spot friction welding of Mg-Mg, Al-Al and Mg-Al alloys", Society of Automotive Engineers, Warrendale (PA), SAE Technical Paper No. 2008-01-0144. 17. Mallick, P.K. and Agarwal, L., 2009, "Fatigue of spot friction welded joints of Mg-Mg, Al-Al and Al-Mg alloys," Society of Automotive Engineers, Warrendale (PA) SAE Technical Paper No. 2009-01-0024. 18. Aota K., Ikeuchi K., "Development of friction stir spot welding using rotating tool without probe and its application to low-carbon steel plates", Welding International, Vol. 23, No. 8, August 2009, pp. 572-580. 19. Ohashi R., Fujimoto M., Mironov S., Sato Y.S., Kokawa H., "Effect of contamination on microstructure in friction stir spot welded DP590 steel", Science and Technology of Welding and Joining, 2009, Vol. 14, No. 3, pp. 221-227. 20. Person N.L., "Tensile-Shear and Fatigue Properties of Resistance and MIG Spot Welds of Some Al Auto Body Sheet Alloys", SAE Paper No. 750463,1975. 21. Chao Y.J., "Ultimate Strength and Failure Mechanism of Resistance Spot Weld Subjected to Tensile, Shear, or Combined Tensile/Shear Loads", Journal of Engineering Materials and Technology, Vol. 125, APRIL 2003, pp. 125-132. 22. Khan M.I., Kuntz ML., Su P., Gerlich A., North T. and Zhou Y., "Resistance and friction stir spot welding of DP600: a comparative study", Science and Technology of Welding and Joining, 2007, Vol. 12, No. 2, pp. 175-182. 23. Karthikeyan R., Balasubramanian V., "Predictions of the optimized friction stir spot welding process parameters for joining AA2024 aluminum alloy using RSM", The International Journal of Advanced Manufacturing Technology, April 2010. 24. Thoppul S.D., Gibson R.F., "Mechanical characterization of spot friction stir welded joints in aluminum alloys by combined experimental/numerical approaches", Materials Characterization, November 2009, Vol. 60, No. 11, pp. 1342-1351. 25. Gean A., Westgate S.A., Kucza J.C., Ehrstrom J.C., "Static and Fatigue Behavior of Spot-Welded 5182-0 Aluminum Alloy Sheet", Welding Journal, March 1999, pp. 80s-86s. 26. Hassanifard S., Zehsaz M., Tohgo K., "The Effects of Electrode Force on the Mechanical Behaviour of Resistance SpotWelded 5083-O Aluminium Alloy Joints", Strain, 2009. 27. Uematsu Y., Tokaji K., "Comparison of fatigue behaviour between resistance spot and friction stir spot welded aluminium
Acknowledgments The authors would like to recognize Richard Osborne, James Quinn, Alan Luo, John Allison, and Robert McCune for their encouragement of this study. This material is based upon work supported by the Department of Energy and the National Energy Technology Laboratory under Award Number No. DE-FC2602OR22910, AUT021 Network of Centers of Excellence, Natural Sciences and Engineering Research Council of Canada (NSERC), Premier's Research Excellence Award (PREA). Such support does not constitute an endorsement by the Department of Energy of the views expressed herein. This paper was prepared as an account of work sponsored by an agency of the United States Government. Neither the United States Government nor any agency thereof, nor any of their employees, makes any warranty, express or implied, or assumes any legal liability or responsibility for the accuracy, completeness, or usefulness of any information, apparatus, product, or process disclosed, or represents that its use would not infringe privately owned rights. Reference herein to any specific commercial product, process, or service by trade name, trademark, manufacturer, or otherwise does not necessarily constitute or imply its endorsement, recommendation, or favoring by the United States Government or any agency thereof. The views and opinions of authors expressed herein do not necessarily state or reflect those of the United States Government or any agency thereof. References 1. K.U. Kainer, Magnesium Alloys and Technology, Wiley-VCH, Cambridge, United Kingdom, 2003. 2. Luo, A.A., "Magnesium: Current and potential automotive applications" JOM. 54 (2): 42-48, 2002. 3. Luo, A.A., Journal of Materials and Manufacturing. SAE Transactions, Warrendale, PA, 411-21, 2005. 4. Friedrich, H.E., and Mordike, B.L., Magnesium Technology— Metallurgy, Design Data, Applications, Springer-Verlag, Berlin, Germany, 2006. 5. S. Begum, D.L. Chen, S. Xu, Alan A. Luo, Met. Mater. Trans. A 39A (2008) 3014 6 C.L. Fan, D.L. Chen, A. A. Luo, Mater Sei. Eng. A 519 (2009) 38 7. J.D. Bernard, J.B. Jordon, MF. Horstemeyer, H. El Kadiri, J. Baird, David Lamb, Alan A. Luo, "Structure-property relations of cyclic damage in a wrought magnesium alloy," Scripta Materialia, 63 (2010) Viewpoint set no. 47, 751-756. 8. T. Pan, 2007, "Friction stir spot welding (FSSW) - a literature review," Society of Automotive Engineers, Warrendale (PA), SAE Technical Paper No. 2007-01-1702. 9. Lin, P.-C, Pan, J, Pan T., 2008, "Failure Modes and Fatigue Life Estimations of Spot Friction Welds in Lap-Shear Specimens of Aluminum 6111-T4 Sheets, Part 1: Welds Made By a Concave Tool," Int. J. Fatigue, 30(1) pp.74-89. 10. Lin, P.-C, Pan, J., Pan, T., 2008, "Failure Modes and Fatigue Life Estimations of Spot Friction Welds in Lap-Shear Specimens
633
alloy sheets", Science and Technology of Welding and Joining, 2009, Vol. 14, No. 1, pp. 62-71. 28. Behravesh B., Liu L., Jahed H., Lambert S., Glinka G., Zhou Y. , "Effect of Nugget Size on Tensile and Fatigue Strength of Spot Welded AZ31 Magnesium Alloy", SAE Technical Paper, 2010, SAE No. 2010-01-0411.
634
Magnesium Technology 2011 Edited by: Wim H. Sillekens, Sean R. Agnew, Neale R. Neelameggham, andSuveen N. Malhaudhu TMS (The Minerals, Metals & Materials Society), 2011
Magnesium Technology 2011 Addendum
Magnesium Technology 2011 Edited by: Wim H. Sillekens, Sean R. Agnew, Neale R. Neelameggham, andSuveen N. Mathaudhu TMS (The Minerals, Metals & Matertals Society), 2011
CONTROLLING THE BIODEGRADATION RATE OF MAGNESIUM-BASED IMPLANTS THROUGH SURFACE NANOCRYSTALLIZATION INDUCED BY CRYOGENIC MACHINING Z. Pu1, D. A. Puleo2, O.W. Dillon, Jr. ', I.S. Jawahir1 'Department of Mechanical Engineering, Center for Manufacturing, University of Kentucky; Lexington, KY 40506, USA 2 Center for Biomédical Engineering, Wenner-Gren Lab, University of Kentucky; Lexington, KY 40506, USA Keywords: biodegradable implants, nanocrystallized grain, cryogenic machining, magnesium alloys gas bubbles generated due to the high corrosion rate impeded further investigation until recently.
Abstract Magnesium alloys are emerging as a new class of biodegradable implant materials for internal bone fixation. They provide good temporary fixation and do not need to be removed after healing occurs, providing the relief to the patients and reducing the healthcare costs. However, premature failure of these implants often occurs due to the high biodégradation rate caused by low corrosion resistance of magnesium alloys in physiological environments. To control biodégradation/corrosion of magnesium alloys, grain refinement on the surface was achieved through machining-induced severe plastic deformation. Liquid nitrogen was used during machining to suppress grain growth. White layers, which consist of nanocrystallized grain structures, are reported herein for the first time in magnesium alloys. By controlling the machining conditions, white layers with various thicknesses were fabricated. In vitro corrosion tests proved that different machining conditions can significantly change the biodégradation rate of magnesium alloys.
In 2005, Witte et al. found results similar to the early researchers through in vivo study of magnesium alloys. A stimulatory effect for bone growth was also reported [4]. The formation of a biomimetic layer comprised of magnesium and calcium phosphate at the implant/bone interface was found to be the cause for accelerated bone formation [5, 6]. Despite their attractive features, little progress has been achieved in controlling the biodégradation rate of magnesium alloys. Alloying and coating are two approaches widely studied [7, 8]. However, alloying may introduce elements which may lead to adverse biological reactions. Stability of the coating under cyclic loading in physiological conditions is a great challenge while the complexity of coating techniques may significantly increase the cost of implant. Mechanical processing of magnesium alloys provides an alternative approach to control the biodégradation rate. Hot rolled magnesium AZ 31 samples were reported to have a marked reduction in biodégradation rate compared with squeeze cast samples [9]. The reduction was attributed to grain refinement from 450 urn to 20 urn. However, further grain refinement by equal channel angular pressing (ECAP) to 2.5 um did not decrease the biodégradation rate. With the same material, Alvarez-Lopez et al. [10] found that samples with 4.5 urn grain size processed by ECAP and followed by rolling had better corrosion resistance than the initial samples with 25.7 urn grain size. Deep rolled magnesium MgCa3.0 samples were reported to have a pronounced reduction in biodégradation rate [11]. Machining with different cutting speeds also leads to different corrosion rates [11].
Introduction In the U.S. alone, physician visits for orthopedic surgery reached 48,066,000 in 2006 [1]. Nine out of the twenty five most common orthopedic surgeries involve repair of bone fractures [2], Internal bonefixationimplants, such as bone plates and screws, are widely used to provide temporary fixation for fractured bones. Stainless steels and titanium alloys are two major biomaterials currently used for these implants. However, their excessively stronger mechanical properties compared to bones may lead to stress shielding. The corrosion and fatigue of these materials will inevitably generate metallic ions and particles that may activate adverse tissue reactions. To avoid further reactions after bone healing, these implants need to be removed during a second surgery, which adds additional morbidity (pain, refracture, etc.) to the patients and increases healthcare costs.
Surface and subsurface integrity in machined products is emerging as the new focus in machining research. The performance of the components can be significantly modified by machining through changes in surface integrity factors, such as microstructure, hardness and residual stresses [12]. Significant grain refinement occurs at the machined surface/sub-surface through severe plastic deformation. Nanocrystallized grains of about 5 - 2 0 nm in size were reported in the white layer of AISI 52100 steel after machining [13]. Ultrafine grains about 175 nm were formed on the machined surface of copper [14]. Due to grain growth caused by the large amount of heat generated during machining, the nanocrystallized grains can be found only at the top surface of the machined component and the cutting speed is limited to very low range. A novel technique based on machining was developed to fabricate thick nanocrystallized layers with the help of the liquid nitrogen cooling [15]. The results proved that
While various approaches are being investigated to increase the bio-inertness of traditional implant materials, magnesium alloys are emerging as a novel biodegradable material in which the relatively fast corrosion phenomenon is used as a unique advantage for temporary fixation implants. The potential of magnesium alloys as a biodegradable implant material was explored by several researchers in the first half of the twentieth century. The results of these research investigations were summarized by Staiger et al. [3]. No systematic reaction occurred and little inflammation was observed in these human trials. A marked stimulatory effect for bone healing was also reported. However, the premature failure of magnesium-based implants due to the poor corrosion resistance in physiological environments and
637
severe plastic deformation under cryogenic conditions can successfully introduce nanocrystallized grain structures to the surface and sub-surface layers. However, only a few studies on controlling the grain refinement through advanced process control have been reported. Also, the relationship between grain refinement, especially in the nanocrystalline range, and biodégradation rates in magnesium alloys is still unknown. Therefore, the aim of the present work was to investigate the microstructural changes under different machining conditions and their influence on biodégradation rate of magnesium alloys incubated in simulated body fluid (SBF). Experimental Work
Figure 2. Edge radius measurement of the cutting tool using ZYGO New View 5300
Work Material
The matrix for the machining experiments is shown in Table I. For each machining experiment, a KISTLER 3-Component Tool Dynamometer was used to measure the cutting forces.
The work material studied was the commercial AZ31 B-H24 magnesium alloy. In vivo tests showed the potential of magnesium AZ31 alloy as a bone implant was significant [16, 17]. The work material was received in the form of 3 mm thick sheet. Disc specimens were made from the sheet and subsequently subjected to orthogonal machining.
No.
Machining Experiments
1 2
The machining experiments were conducted on a Mazak Quick Turn-10 Turning Center equipped with an Air Products liquid nitrogen delivery system. The experimental setup is shown in Figure 1.
3 4
Table I. Matrix for the machining experiments Cutting Feed Cooling Method Edge Radius (um) Rate Speed (m/min) (mm/rev) Dry 30 100 0.1 100 0.1 Cryogenic 30 100 0.1 Cryogenic 68 Cryogenic
74
100
0.1
In vitro Corrosion Test To mimic the human body environment, a simulated body fluid (SBF) was prepared: 8.0 g/1 NaCl, 0.4 g/1 KC1, 0.14 g/1 CaCl2, 0.35 g/1 NaHC03, 1.0 g/1 C6H1206 (D-glucose), 0.2 g/1 MgS04-7H20, 0.1 g/1 KH2P04- H20 and 0.06 g/1 Na2HP04-7H20. The pH of the SBF was adjusted to 7.4. The solution was kept in an incubator to maintain the temperature at 37 ± 1 °C. To evaluate the biodégradation rates, hydrogen evolution method [18] was used to continually monitor the corrosion process for 7 days. To reduce the effects of pH increase and accumulation of corrosion products on corrosion rate, large solution volume/surface area (SV/SA) ratio (SV/SA=433) was used [19]. 10 mL graduated cylinder (0.1 mL interval) Incubator
Figure 1. Machining setup with an Air Products liquid nitrogen delivery system
Simulated body fluid Machined sample
The machining conditions controlled during the experiments were cooling methods and the edge radius of the cutting tool. For dry machining, no coolant was used. For cryogenic machining, liquid nitrogen was applied to the machined surface from the clearance side of the cutting tool. The cutting tools used were uncoated carbide C5/C6 inserts from Kennametal. These cutting tools were ground to three different edge radii. The actual edge radius before machining was measured using a ZYGO New View 5300 measurement system which was based on white light interferometry. A sample measurement is shown in Figure 2.
Hydrogen gas bubble
Figure 3. In vitro corrosion test setup Characterization Method Metallurgical samples were cut from the machined discs. After cold mounting, grinding and polishing, acetic picric solution was used as the etchant to reveal the grain structure. Optical and scanning electron microscopes were used to observe the
638
The influence of machining conditions on white layer formation is clearly seen in Figure 6. Dry machining of magnesium alloys using a tool with 30 um edge radius did not lead to white layer formation. However, using the same edge radius, a white layer of about 7 um thickness was formed under cryogenic machining conditions. Under the same cooling condition, the edge radius of the tool played a remarkable role in white layer formation. The thickness of the white layer was increased to about 15 urn when the edge radius was increased to 68 urn. However, further increase in edge radius to 74 irm reduced the white layer thickness.
microstructure of the magnesium alloys. An atomic force microscope (AFM) was used to explore the possible structure of the top layer of the machined samples. For AFM characterization, the samples were observed after grinding and polishing but without etching. The chemical composition of the top layer was determined by energy dispersive spectroscopy (EDS). Results and Discussion Microstructure The initial microstructure before machining experiment is shown in Figure 4. The grain boundaries are clearly visible throughout the sample.
Figure 6. White layer thickness under different machining conditions
Figure 4. Initial microstructure before machining experiment
Force Analysis Cutting force and radial force data during machining were analyzed to explore the influencing factors in white layer formation. Both the cutting force and radial force were stable during the 30 second cutting time, indicating little tool-wear and tool/chip adhesion. The cutting forces for all the four experiments remained at about 180 N. However, a significant influence of tool edge radius and cooling method on radial force was present. Figure 7 shows the radial force recorded for the machining experiments. In the cryogenic group, the radial force became larger with increasing edge radius of the tool. A large increase was observed when the edge radius was increased from 30 um to 68 pm, which corresponds to the large increase in white layer thickness. The radial force was further increased using the tool with 74 (im edge radius. Higher stresses introduced by the large edge radius lead to more deformation twinning within the grains compared with 30 pm and 68 urn. However, the white layer thickness decreased. This may indicate that white layer formation was dependent on both mechanical and thermal effects, which agrees with the research in white layer formation in steels [20].
Figure 5. Microstnicture of magnesium alloy discs after machining: (a) dry machining, edge radius = 30 urn, (b) cryogenic machining, edge radius = 30 um, (c) cryogenic machining, edge radius = 68 urn, (d) cryogenic machining, edge radius = 74 um. The microstructures of the samples machined under different conditions are presented in Figure 5. Although the grain structures of the bulk material were similar under all machining conditions, significant differences were apparent in the surface layers of different machined samples. For dry machined samples (Figure 5(a)), grain boundaries are clearly visible on the surface. For cryogenic machined samples (Figure 5(b), (c) and (d)), surface layers of different thicknesses, where grain boundaries are not discernable, were formed. This layer of indiscernible grain structure was also reported in other materials after machining, especially in steels, where the term "white layer" was frequently used [13]. While significant research has been done in white layer formation in steels, white layer in magnesium alloys is reported here for thefirsttime.
Figure 7. Radial force measurement under different machining conditions
639
EDS Analysis
AFM Characterization
To explore the chemical composition of the white layer, energy dispersive spectroscopy (EDS) was used. Figure 8 shows the results of the EDS analysis. Only a little chemical difference between the bulk material and the white layer was found.
The ability of atomic force microscope (AFM) to measure grain size was studied by several researchers [21, 22]. Phase imaging tapping-mode AFM was reported to successfully provide grain boundary details and the accuracy of the grain size measurements was comparable to TEM measurement with ±10% depending on AFM calibration accuracy. After grinding and polishing, the top portion of the white layer on Sample No. 3 was observed using tapping-mode AFM. The phase image is shown in Figure 10. While no structures were discernable in optical microscope or SEM pictures, some grain-like features were present in the AFM phase image. The mean size of the features was about 45 nm and all the features were smaller than 100 nm. Based on the AFM picture and the literature review on nanocrystallized grains in white layers of steels and copper, a preliminary conclusion can be made that the white layer on the machined surface of magnesium alloys consisted of nanocrystallized grain structures. Further investigation using transmission electron microscope (TEM) will be conducted to verify this conclusion.
Figure 8. EDS analysis of the bulk material and the white layer SEM Characterization Figure 10. AFM tapping mode phase image of the white layer
Nanocrystallized grain structures were found in white layers in steels [13]. A thick nanocrystallized layer was also found on copper processed by a surface mechanical grinding treatment (SMGT) under liquid nitrogen cooling [15]. Due to the similarity that both cryogenic SMGT and cryogenic machining in producing severe plastic deformation, it is expected that the white layer formed on cryogenically machined magnesium AZ31 samples consisted of nanocrystallized grain structures. To explore this assumption, scanning electron microscope (SEM) and atomic force microscope (AFM) were used to observe the white layer formed on Sample No. 3 (Table I).
In vitro Corrosion Test The influence of different machining conditions on the biodégradation rate of magnesium alloys was investigated by the in vitro corrosion tests. Figure 11 shows accumulative hydrogen evolution per unit area from magnesium AZ31 discs machined under different conditions. The machined samples were left in air for two weeks before the in vitro corrosion test, which simulated the passivation stage of implant preparation; a passivated layer was formed on the machined surface due to the oxidation of magnesium in air. This passivated layer was attributed to the different shape of the biodégradation curves compared with other studies, where a very high biodégradation rate occurred at the beginning of the corrosion test.
Figure 9(a) shows that grain boundaries are clear except the top layer of the sample, and twinning was present more than 100 um away from the machined surface. Figure 9b shows that grain boundaries in the white layer were not visible even under x5000 magnification.
Preliminary corrosion test results clearly demonstrated the influence of machining conditions on corrosion resistance of magnesium alloys in physiological environment. The slowest corrosion rate occurred on the machined sample with the thickest white layer. However, the sequence of corrosion rates did not agree with the sequence of white layer thickness. Other factors, like deformation twins, may also contribute to the variation in corrosion rate. Other corrosion analyses, such as weight loss and electrochemical methods, will be conducted to verify the results from this preliminary experiment.
Figure 9. SEM pictures of Sample No. 3: (a) x500 (b) x5000.
640
biodégradation rate customized to individual medical demands can be manufactured. References 1. D.K. Cherry, E. Hing, D.A. Woodwell, E.A. Rechtsteiner, "National Ambulatory Medical Care Survey: 2006 Summary" (National Health Statistics Reports, Number 3, August 6, 2008). 2. W.E. Garrett, Jr., M.F. Swiontkowski, J.N. Weinstein, J. Callaghan, R.N. Rosier, D.J. Berry, J. Harrast, G.P. Derosa, "American Board of Orthopaedic Surgery Practice of the Orthopedic Surgeon: Part-II, Certification Examination Case Mix", The Journal of Bone and Joint Surgery, 2006, no. 88:660667.
Figure 11. Hydrogen evolution during in vitro corrosion test An interesting phenomenon during the corrosion test was found between the cryogenic machined sample using tool with 68 urn edge radius and the dry machined sample using tool with 30 um edge radius. The former had the thickest white layer (about 15 um), while the latter did not have white layer. During the corrosion process, relatively large black spots caused by pitting were formed on the sample without white layer. However, for the sample with the thickest white layer, the whole surface gradually turned dark over time without formation of large black spots. This difference suggested that the sample machined under cryogenic condition underwent more homogeneous corrosion than the one under dry condition. This homogeneity is expected to be beneficial in orthopedic implant applications. The most attractive benefit is the formation of a uniform implant/tissue reaction layer, which may lead to better osseointegration. Also, stress concentration may be avoided due to the homogeneity of the pitting.
3. Mark P. Staiger, Alexis M. Pietak, Jerawala Huadma, George Dias, "Magnesium and its alloys as orthopedic biomaterials: A review", Biomaterials, 2006, no. 27: 1728-1734. 4. Witte F, Kaese V, Haferkamp H, Switzer E, Meyer-Lindenberg A,Wirth CJ, H. Windhagen, "In vivo corrosion of four magnesium alloys and the associated bone response", Biomaterials, 2005, no.26:3557-3563. 5. Liping Xu, Guoning Yu, Erlin Zhang, Feng Pan, Ke Yang, "In vivo corrosion behavior of Mg-Mn-Zn alloy for bone implant application", Journal of Biomédical Materials Research Part A , 2007, 83A: 703-711. 6. Zhang E, Xu L, Yu G, Pan F, Yang K, "In vivo evaluation of biodegradable magnesium alloy bone implant in the first 6 months implantation", Journal of Biomédical Materials Research Part A, 2008,90A:882 - 893.
Conclusion The present study shows that the biodégradation rate of the magnesium alloys can be effectively altered by controlling the machining conditions. White layer in magnesium alloys is reported for the first time. The slowest corrosion rate occurred on the cryogenically machined sample with the thickest white layer. Also, corrosion on this sample was found to be more homogeneous compared with the dry machined sample that did not have a white layer.
7. Guangling Song, "Control of biodégradation of biocompatible magnesium alloys", Corrosion Science, 2007, no. 49: 1696-1701. 8. Cuilian Wen, Shaokang Guan, Li Peng, Chenxing Ren, Xiang Wang, Zhonghua Hu, "Characterization and degradation behavior of AZ31 alloy surface modified by bone-like hydroxyapatite for implant applications", Applied Surface Science, 2009, no.255: 6433-6438.
The AFM phase image suggests that the white layer consisted of nanocrystallized grain structures. The remarkable grain refinement was the combined result of dynamic recrystallization through machining-induced severe plastic deformation and effective suppression of grain growth by liquid nitrogen cooling.
9. H. Wang, Y. Estrin, Z. Zuberova, "Bio-corrosion of a magnesium alloy with different processing histories", Materials Letters, 2008, no.62: 2476-2479.
The edge radius of the cutting tool has an important influence on the thickness of the white layer. A combined thermal-mechanical effect for white layer formation was detected.
10. M. Alvarez-Lopez, Maria Dolores Pereda, J.A. del Valle, M. Fernandez-Lorenzo, M.C. Garcia-Alonso, O.A. Ruano and M.L. Escudero, "Corrosion behaviour of AZ31 magnesium alloy with different grain sizes in simulated biological fluids", 2009, Ada Biomaterialia. Article in Press, Corrected Proof.
The preliminary results from the present study reveal a great opportunity to control biodégradation rate of magnesium alloys through advanced process control techniques. Existing knowledge on surface integrity of machined products and various predictive modeling techniques can significantly facilitate the development of a machining-based process to control the biodégradation rate of magnesium alloys. In vitro corrosion tests can correlate the surface integrity factors with the corresponding biodégradation rate. In the end, magnesium-based implants with specific
11. B. Denkena, A. Lucas, "Biocompatible Magnesium Alloys as Absorbable Implant Materials -Adjusted Surface and Subsurface Properties by Machining Processes", 2007, Annals of the CIRP, no.56: 113-116. 12. R. M'Saoubi, J.C. Outeiro, H. Chandrasekaran, O.W. Dillon Jr., I.S. Jawahir, "A review of surface integrity in machining and its impact on functional performance and life of machined
641
products", International Journal of Sustainable Manufacturing , 2008, no. 1:203-236. 13. A. Ramesh, S.N. Melkote, L.F. Allard, L. Riester, T.R. Watkins, "Analysis of white layers formed in hard turning of AISI 52100 steel", Materials Science and Engineering A, 2005, no.390: 88-97. 14. R. Calistes, S. Swaminathan, T.G. Murthy, C. Huang, C. Saldana, M.R. Shankar and S. Chandrasekar, "Controlling gradation of surface strains and nanostructuring by large-strain machining", Scripta Materialia, 2009, no.60:17-20. 15. W.L. Li, N.R. Tao, K. Lu, "Fabrication of a gradient nanomicro-structured surface layer on bulk copper by means of a surface mechanical grinding treatment", Scrivta Materialia, 2008, no.59: 546-549. 16. Yaohua He, Hairong Tao, Yan Zhang, Yao Jiang, Shaoxiang Zhang, Changli Zhao, Jianan Li, Beilei Zhang, Yang Song and Xiaonong Zhang, "Biocompatibility of bio-Mg-Zn alloy within bone with heart, liver, kidney and spleen", Chinese Science Bulletin, 2009, no. 54: 484-491. 17. Duygulu 0,Kaya RA,Oktay G.Kaya AA, "Investigation on the potential of magnesium alloy AZ31 as a bone implant", Material Science Forum, 2007, no. 546-549: 421-424. 18. Song, G., Atrens, A., St John, D. H., "An hydrogen evolution method for the estimation of the corrosion rate of magnesium alloys", Proceeding of Magnesium Technology 2001, TMS Annual Meeting. New Orleans, LA. February 11-15, 2001. 19. Lei Yang, Erlin Zhang, "Biocorrosion behavior of magnesium alloy in different simulated fluids for biomédical application", Materials Science and Engineering C, 2009, no. 29:1691-1696. 20. Sangil Han, "Mechanisms and Modeling of White Layer Formation in Orthogonal Machining of Hardened and Unhardened Steels" (Ph.D. thesis, Georgia Institute of Technology, 2006) 21. C. H. Pang, P. Hing, A. See, "Application of phase-imaging tapping-mode atomic-force microscopy to investigate the grain growth and surface morphology of TiSi2", The Journal of Vacuum Science and Technology B, 2002, no. 20: 1866-1869. 22. Alexander Luce, "Atomic Force Microscopy Grain Structure Characterization of Perpendicular Magnetic Recording Media" (2007 REU Research Accomplishments, National Nanotechnology Infrastructure Network, pp. 134-135).
Magnesium Technology 2011 Edited by: Wim H. Sillekens, Sean R. Agnew, Neale R. Neelameggham, andSuveen N. Malhaudhu TMS (The Minerals, Metals & Materials Society), 2011
AUTHOR INDEX Magnesium Technology 2011 A
Agnew, S Alam, M Alba-Baena,N Alderman, M Aljarrah, M Amini, S Amoorezaei, M Anasori, B Anderson, W Antonyraj, A Atiya, G
B
Badarinarayan, H Bae,J Baird,J Bamberger, M Bang, W Bao, Y Barnett, M Barrett, C Barsoum, M Bart, F Beckermann, C Behravesh, S Berkmortel, R Berman,T Bermudez, K Bernard, J Bettles, C Bhatia, M Bhole, S Bichler, L Blachere,A Blawert, C Boehlert,C Bohlen, J Boismier, D Brar,H Brennan, S Brown, R Burke,P
C
Campos, R Carlson, K Carsley,J Cavin,0 Chang, Y Chen.C Chen.D Chen,0 Chen,Z Cheng, Y Cho,K Choi, H Choo, D
D
D'Errico, F Dai,J Danzy, J
Darling, K Das, S Davis, B Decker, R DeLorme.R Deng,C Dharmendra, C Dillon, 0 Ding, W Doherty, R Donlon, W Du,J
379 553 443 187 565 463 101 463 61 583 249
E
Easton, M El-Kaddah, N El Kadiri, H Elsayed, A Esen,Z Essadiqi, E
629 261 301,313 249 385 543 289 295 463 435 93 629 93 599 549 67 227 325 629 79 435 507 85 113, 373 413 401 549 7 481
F
Fabry, B Fang,X Farè, S Farzadfar, S Feng,N Feyerabend, F Firrao, D Fletcher, M Foley, D Friedman, P Frizon, F Fu, P
G
Gao, B Gao,Y Garces, G Geng,J Gharghouri, M Gibson, J Gibson, M Grantham, J Grassini, S Gratz, E Groëbner, J Gupta, A Gupta, M Gurevich, S
507 93 389 187 267,385 73 629 217 85 617 345 443, 447 143, 147
H
Haddad, D Hamada,G Hammi,Y Hamouda, A Hartwig, K Hashimoto, A Hector, L Hicks,A Higashida,K
19 181 501, 605
643
453 49 187, 345, 379 85,599 345,379 413 169 513, 637 157, 161, 181 389 599 537
167 119 285,295, 301, 313, 583 475 457 339, 565
409 227 19 339 43 17 493 79 559 395 435 161
35 73 19 289 595 55 167, 227 629 493 39 167 363 553 101
611,617 369 583 553 559 369 13 501,605 273
Höche,D Homayonifar, M Homma, T Honjo,T Hono, K Horstemeyer, M Hort,N Hotchkiss, A Howe,J Hu, H Hu, W Hu, X Huang, J Huang, Y Huber,N Hutchinson, C
I
Inoue, T.
J
Jahed,H Jain,V Janecek, M Javaid, A Jawahir, 1 Jayaraj, J Jekl,J Jiang, L Jiang, Y Jin.L Jonas, J Jones, J Jones, M Jones, T Jordon, J Jung, 1
K
Kainer,K Kalidindi,S Kamado, S Kang,J Kang, X Katsman, A Kawamura, Y Kecskes, L Keselowsky, B Keyvani, M Kim,G Kim, H Kim, I Kim,J Kim, M Kim,N Kim, S Kim,W Kim, Y Kipouros, G Klassen, R Kondoh, K Konishi, H Kou, S Kozlov,A Krajewski, P Kuji,T
507 321 223 431 223, 245, 261 55, 67, 285, 301, 313, 349, 357, 501, 583, 605, 629 17, 113, 125, 169 413 85 137, 469 43 35 85,599 125 321 227
Kulkarni,N Kuo, H
L
Lakes, R Lambert, S Lambertin, D Larsen, S Le Beau, S Lee, G Lee, H Lee, S Letzig, D Levinson, A Li,B Li,J Li,M Li,W Li,X Li,Z Lilleodden, E Liu, H Liu,Z Lu, C Lugo, M Luo, A Luo,R
..211
629 565 589 565 513, 637 245 93 333 543 363 333 61,217, 599 443 425 55, 67, 357, 629 49, 339
M
Ma,Q Mahmudi, R Manavbasi, A Manuel, M Mao,P Marin, E Martin, H Mathaudhu, S Matsuzaki, K Matteis, P Mayama, T McCune, R Mendis, C Meratian, M Milshtein, J Mironov, S Mirshahi, F Mishra, R Miwa, K Moitra, A Mosler,J Mu, W Mukai,T Murakami, Y Murakoshi, Y Muralidharan, G Muth, T
5, 113, 125, 169,373,507 389 195,223 307 161 249 229, 273 453 401 571 321 385, 589 143 143, 147, 151,229 151 261 131, 285, 349 151 49 481 79 425, 475 447 443 167 395 431
N
Nakawaki, S Nayyeri, G Nebebe, M Nguyen, Q Nibhanupudi, S Nie,J Nimityongskul, S Niu,X
644
549 523
443 629 435 413 85 151 385 595 373 389 295 255 107 623 443,447 161,537 321 623 125 157 357 161,267 537
301,583 571 519 401 125 583 501,605 453 485 493 273 531 245,261 577 39 199 577 205, 307, 333, 363, 389, 565 107 325 321 537 25,211,239 107 485 187 187
223 571 321 553 519 167 443 137
o
Oh-ishi, K Ohashi.T Ohkubo,T Okamoto, K Omura,N Oppedal.A Osawa, Y
P
Pal, U Paliwal, M Park,J Park, W Patel, H Pati, S Peng, L Peter, W Petersen, E Petit, C Polesak, F Pollock, T Powell, A Prasad, Y Presser, V Provatas, N Pu,Z Puleo, D Punessen, W
R
Radhakrishnan, R Randman, D Rao,K Reck,J Rettberg, L Rimkus,N Ringer, S Robson.J Root,J Rossetto, M
Shook, S Shyam, A Sikand,R Sillekens, W Singh, A Singh, J Sivilotti, 0 Slade, S Sohn, Y Solanki, K Somekawa,H Song, G Song, P Song, S Staiger, M Stanford, N Stinson, J Strâsky,J Stutz, L Su,J Su, X Sun, M Sun,Z Suresh,K
223, 245 273 245 73, 199 107 301,313 239
39 49 151 143, 147 629 39 161 187 413 481 379 217, 599 39 169 463 101 513,637 513,637 113
T
Tada, S Tadano,Y Takahashi, N Tamura, T Tang,T TerBush,J Tome, C Tong, L Trinkle, D Tschopp, M
413 187 169 523 61 119 255 289 595 493
U
Umeda,J Utsunomiya, H
V S
Sachdev, A Saha, P Sahoo, M Sakai,T Sakuragi.K Salgado-Ordorica, M Samson, H Sanjari, M Sano,T Samtinoranont, M Sasaki, M Sato, M Sato, Y Scavino, G Scheuermann, T Schmid-Fetzer, R Schumacher, P Sediako.A Sediako, D Sha, G Shen,W Shi,Z Shimizu, T
Van Lieshout, K Verma, R Virtanen, S Viswanathan, S Vogel, S Vrâtnâ,J
363 175, 559 443 369 431 113 553 339 345 401 73 431 199 493 413 167 255 233 79,233 255 531 35 485
W
Walton, C Wang, P Wang,Q Wang, Y Wang, Z Waterman, J Watkins,T Weaver, M Weber, J Werkhoven, R Wilkinson, D Witte, F Wood,T Wu, G Wu, K
645
233 85 363 419 25, 211, 239 395 31 3 549 325 25,211,239 513,611, 617 195 531 403 289 413 589 373 565 629 181 137 169
107 279 73 107 349 217 313 195 13 295,357
475 369
419 565 409 175, 559 233,313 589
501,605 349, 501, 583, 605 73 43,523,531 35,43 403 187 119 413 419 307 17 443 181 195
X
Xi,Z Xu, S Xu,Z
Y
Yakoubi, S Yang, F Yang.K Yang,Q Yang, S Yang, Y Yasi,J Yi,S Yin, D Yoon, E Yoon, U Yu,J Yuan, G Yuan, W Yue,S
z
Zahrani, E Zahrani, M Zhang, C Zhang, Q Zhang, S Zheng, M Zheng, Y Zhou, L Zhu, L Zhu, S Zhu,T
537 195 611
481 35 543 199 35, 513 629 13 113 73 589 151 345 157 205 339
577 577 267 469 157 195 399 125 623 167 161
646
Magnesium Technology 2011 Edited by: Wim H. Sillekens, Sean R. Agnew, Neale R. Neelameggham, andSuveen N. Malhaudhu TMS (The Minerals, Metals & Materials Society), 2011
SUBJECT INDEX Magnesium Technology 2011 3
3D Microstructure
A
Adhesion AE44 Age Hardening Ageing Aging Alumina Leaching Aluminum Anisotropy Anisotropy of Mechanical Properties Annealing Anodic Oxidation Armor Asymmetric Rolling Atom Probe Tomography Atomic Force Microscopy Auger AZ31 AZ31 Alloy AZ41 AZ51 AZ61 AZ61 Magnesium Alloy AZ91D
B
Ballistic Performance Bioabsorbable Biocompatibility Biodegradable Implants Biodegradable Magnesium Implants Biodegradation Biomédical Implants Biomédical Magnesium Alloys Bismuth
c
Caliber Rolling Calphad CaO Ca02Al 2 0 3 Casting Casting Defect Cathodic Electrophoretic Deposition Cavitation Chromium C0 2 Reduction Coating Coatings Composites Compression Test Compressive Loading Compressive Strength Computational Material Design Containerless Melting Continuous Casting Conversion Coating Cooling Curve Cooling Rate Corrosion Corrosion Resistance Crack Propagation
Creep Creep Resistance Creep Strain Cryogenic Burnishing Cryogenic Machining Crystal Plasticity Crystallographic Structure Cyclic Stress-Strain
583
543 493 255 239 245 43 325,549 205,475 199 49 623 425 187 223,255 617 507 369,389,395,553 589 553 553 599 67 523
D
Damage Evolution Deformation Deformation Twinning Dendrite Density Functional Theory Design Die-Cast AZ91 Diecasting Diffusion Directional Solidification Dislocations Drawing Ductility Dynamic Recrystallization Dynamic Strain Aging
E
EBSD ECAE Electrochemistry Electrolysis Electrolyte Electromagnetic Elemental Partitioning EMC EMS Encapsulation Engine Pulley Environment Equal Channel Angular Pressing Extrusion
425 401,413 409 399,403,637 17 409 419 399 577
211 167 131 43 93, 119, 167 107 543 385 519 19 403,537 531 457 339 273 169, 239 161 119 151 519, 523 35 577 403,409,513,519,531,611 537, 543 349
F
Faceted Particle Fatigue Fatigue Modeling Fatigue Strength Finite Element Analysis Forging Formability Forming Limit Fraction Solid Fractography Fracture Fracture Mechanism Fracture Toughness Friction Stir Processing Friction Stir Spot Welds
G
G.P.Zone Galvanic Corrosion Global Situation Grain Refinement
647
79, 85,217,223,227 571 73 513 637 273, 279, 313, 321 249 61
357 345 279, 313 101 13 401 61 131 549 101, 113 13 419 131,447 373, 379 333, 379
301,389,475 559 435 31 35 151 217 151 151 435 151 7 195, 419 239,419,443
175 55,67,629 67 73 273 485 187, 373, 395,485 321, 373 107 61 119 25 85 199,205, 565 629
245 493 5 175, 181,211,513, 559
Grain Refining Gravity Casting
H
Hardening HCP Metal Heat Index Heat Transfer Coefficient Heat Treatment Heavy Rolling Heterogeneous Deformation Hexavalent High Pressure Torsion High Speed Rolling High Temperature Creep Homogenization Hot Extrusion Hot Rolling Hot Tearing Susceptibility Hot Tears Hot Torsion Hot Workability Hot-extrusion Hydrogen Storage Materials Hydrostatic Pressure Hysteresis Loops
I
I Phase In-Situ In-vivo Corrosion Incremental Step Test Injection Speed Intergranular Corrosion Intermetallic Phase Intermetallics Inverse Method Isothermal Heat Treatments
K
Kinetic Analysis.
L
Lightweight Alloys Liquid Melt Infiltration Liquidus Temperature Long Period Stacking Ordered (LPSO) Structure Low-cycle Fatigue LPSO
Magnesium Corrosion 507 Magnesium Development 5 Magnesium Market 3 Magnesium Powder Metallurgy 481 Magnesium Powders 481 Magnesium Production 7 Magnesium Research 5 Magnesium Sheet 143, 373 Magnesium Single Crystal 349 Magnesium-Aluminum Alloy 431 Magnesium-Rare Earth 565 Magnesium-Zirconium Alloy 435 Magnetron Sputtering 617 Mechanical Alloying 431 Mechanical Couplings 493 Mechanical Properties 161, 195,229,457, 553, 571, 599 Membrane Stability 39 Mg 611 Mg Alloy 199,229 Mg Alloy Chip 485 Mg Alloys 131 Mg-5Sn Alloy 571 Mg-Al Alloy 447 Mg-Al Binary Alloys 49 Mg-Matrix Composite 463 Mg-Nd-Zn-Zr Alloys 249 Mg-Nd-Zn-Zr-Gd/Y Alloys 255 Mg-Si Alloy 577 Mg-Sn-Ca-Al-Si Alloy 169 Mg-Ti Thin Films 617 Mg-Zn-Ce 267 Mg-Zn-Gd Alloys 157 Mg-Zn-Ca-Mn Alloy 195 Micro-Alloying 227 Micro-Arc Oxidation 537 Microcompression 321 Microhardness Evolution 589 Microstructure... 61, 107, 113, 119, 161, 195,261,345,357,363,571 Microstructures 599 Modeling 93,583 Modeling of Corrosion 605 Modification 577 Modified Oxidation Film 623 Molecular Dynamics 295, 325 Molecular Dynamics Simulation 349 Molten Salt 35 Multiphase Diffusion 49 Multi polar 31
447 73
249 279 205 137 73,333 369 273 519 589 369 233 379 363,475 565 125 93 379 169 485 431 385 61
157 345 17 61 107 501, 605 447 357 137 167
.169
85 463 35 157 61 229
N
Nano-Crystalline Materials Nano-indentation Nanocomposite Nanocomposites Nanocrystalline Nanocrystallized Grain Nanoindentation Necking Neutron Diffraction Non-SF6 North America Nucleation
M
Magnesium 25,31,39,43,85, 101, 107,175, 187,211,233,285, 289, 295, 301, 313, 321, 325, 339, 345, 357, 385, 395, 403,409, 413, 443, 453, 457, 519, 531, 549, 559, 583, 595 Magnesium Alloy... 125, 137, 181,223,307,333,363,401,501,537, 543, 553,605,623 Magnesium Alloy Development 161 Magnesium Alloys... 13, 19, 55, 79,93,113,239, 245, 267,425, 513, 629,637 Magnesium Alloy Sheet 147 Magnesium AZ31-B Alloy 119 Magnesium Billet 151 Magnesium Coil 143
o
648
431 79 447 443, 463 453 637 617 307 595 131 3 285
Orientation Relationship
249
Particle Size Distribution
175
Phase Diagram Phase Identification Physical Vapor Deposition Pitting Corrosion Plasma Electrolytic Oxidation Plastic Deformation Plasticity Porosity Powder Metallurgy Precipitate Precipitates Precipitation Precipitation Hardening Precipitation Sequence Processing Maps Pure Magnesium
Q
Quasicrystal.
R
R-value Rapid Solidification Process Rare Earth Rare Earth - Lanthanum Rare Earths Recrystallization Recycling Reduction Replica Resistance Spot Weld Roller Hemming Rolling
S
Segregation Semi-Solid Shear Bands Shear Strain Sheet Shrinkage Porosity Shuffling Single Crystal Sintering Phenomena Si0 2 -Al 2 0 3 Sol Solidification Solute Cerium Solute Strengthening SOM Spinning Water Atomization Process (SWAP) Sputtering Squeeze Casting Stent Stents Strain Rate Strain-Life Strength Strip Casting Structure-property Relations Superplasticity Surface Treatment Sustainability
T
TEM..
Temperature Effects 349 Tensile Behavior 205 Tensile Properties 577 Tensile Strength 73,239 Texture 187, 199,211,233, 301, 339, 363, 373, 565,583, 599 Texture Evolution 369 Thermal Stability 453 Thermodynamics 339 Thermomechanical Processing 599 Thixomolding 599 Ti 611 Toughness 25 Transient 101 Transmission Electron Microscopy 223 Trivalent 519 Twin Accommodation Effects 313 Twin Nucleation 295, 325 Twin Roll Cast 147 Twin Roll Casting 143,261 Twinning 273,285,289,301, 321, 389, 595 Twins 25
267 147 611 501, 605 543 13,307, 595 301,583 61 457,463,475 255 289 217,227,267 261 249 169 279
..239
u
307 425 167 507 113 369 7,485 39 67 629 389 369
Ultrafine Microstructure Ultrasonic Dispersion
Vacuum Aluminothemic Reduction
w
Warm Forming WE43 Work Hardening
147 107 389 211 395 93 285 295, 325 481 623 267,559 333 217 39 425 611 137 413 419 349 61 131,453 143 55 385 403 19
X
X-ray Diffraction XPS
Y
Yttrium..
Zirconia Zirconium (Zr)
175,255,289
649
19 443
43
395 379 205
617 507
.339
523 175, 181