Deformation inhomogeneity in large-grained AA5754 sheets
Guozhen Zhu,Xiaohua Hu,Jidong Kang,Raja K. Mishra,David S. Wilkinson
DOI: https://doi.org/10.1016/j.msea.2011.02.017
2011-01-01
Abstract:Research highlights ► Microstructure and strain relationship at individual grain level was studied. ► ‘Hot spots’ nucleate early and most keep growing throughout deformation stages. ► ‘Hot spots’ are correlated with ‘soft’ grains and soft-evolution grains. ► Grains with high Schmid factors tend to be ‘soft’ grains. ► Grains with the 〈1 0 1〉 direction close to tensile axis tend to become softer. Abstract Models for deformation and strain localization in polycrystals that incorporate microstructural features including particles are computationally intensive due to the large variation in scale in going from particles to grains to a specimen. As a result such models are generally 2-D in nature. This is an issue for experimental validation. We have therefore studied deformation heterogeneities and strain localization behavior of coarse-grained alloys with only two grains across the sample thickness, therefore mimicking 2-D behavior. Aluminum alloy sheets (AA5754) have been investigated by a number of surface techniques, including digital image correlation, slip trace analysis and electron backscattered diffraction, at the individual grain level. Local strain concentration zones appear from the very beginning of deformation, which then maintain sustained growth and lead, in one of these regions, to localization and final fracture. These ‘hot spots’ occur in areas with locally soft grains (i.e. grains with 〈0 0 1〉 or 〈1 0 1〉 close to the tensile direction) and soft-evolution orientations (i.e. grains with 〈1 0 1〉 close to the tensile direction). These grains can be correlated with Taylor and/or Schmid factors. Keywords Strain concentration AA5754 Coarse-grained Digital image correlation 1 Introduction There is growing interest in aluminum alloys within the automotive industry due to their high strength and stiffness to weight ratio, good formability, corrosion resistance, recycling potential and so on. Non-heat treatable Al–Mg sheets (5xxx series) find applications for body-in-white (BIW) internal structural parts and occasionally for outer panels. [1,2] However, raw material and manufacturing costs impede greater utilization of this material. Continuous strip casting (CC) provides a method for fabricating aluminum sheets at lower cost than conventional methods involving direct-chill casting (DC). However, the limited thermo-mechanical processing of CC material, a direct consequence of the low as-cast thickness limits the development of a homogeneous microstructure and further affects the formability of these alloys. [3,4] The formability of Al-Mg sheets is limited by plastic strain localization. The material constitutive behavior is one of the major factors determining the onset of strain localization, in terms of both diffuse and localized necking. Of equal importance, the various microstructural constituents can provide sources of deformation inhomogeneity which can in turn cause premature localized deformation and failure. The understanding of the role of various microstructure factors, i.e. grain structure, on strain localization is still incomplete. The crystallographic texture (e.g. the distribution of crystal orientations (ODF)) is a major cause of deformation anisotropy which can lead to problems such as earring and surface roughening. Earlier experiments on polycrystalline materials [5–7] have demonstrated the non-crystallographic nature of shear bands. However, these results, based on the incomplete information of the overall microstructure, cannot fully describe the effect of grain orientation on the path of propagation of the localization necking bands. To date, both characterization and simulation of the deformation process in a full three-dimensional microstructure with features at different scales ranging from small second phase hard particles to grains up to the much larger specimen scale are still not practical. Simulations of deformation using model microstructures [8,9] have shown that shear band formation is not only sensitive to the variation of statistical texture but also the spatial grain orientation distributions, e.g. the alignment of texture colonies, where a texture colony consists of a cluster of grains with similar grain orientations of a specific texture component. Similarly, with the assumption that there is no substantial change of grain orientations prior to macroscopic strain localization (i.e. localized necking), Hu et al. [10] have represented grain-level inhomogeneities using a constant relative Taylor factors of grains during uniaxial tensile deformation. Strain localization is found to be initiated in soft grains (with small Taylor factors) and propagate through a ‘soft path’ (the arrangement of grains of low Taylor factors) to form a macroscopic localized necking band. Experiments on oligocrystals (i.e. samples with a few grains of millimeter size) have magnified grain-level inhomogeneities to a level that enables more complete characterization and provides a bridge linking to microstructure-based simulation, which for computational reasons are generally two dimensional in nature. In contrast with other strain measurement techniques e.g. synchrotron X-ray diffraction [11] ), digital image correlation (DIC) provides a practical method that can be widely applied to detect local displacement out to large strain with high accuracy [12] . Combined with the surface microstructure characterization techniques (i.e. EBSD), deformation inhomogeneity in large grains can be directly correlated to the microstructure and its evolution. Raabe et al. [13] undertook detailed strain mapping under compression on oligocrystals and proved that macroscopic boundary conditions and grain interactions work together to develop deformation heterogeneities. They also introduced the kinematical Taylor factors to describe the relative hardness of grains and relate them to variation of local deformations between grains. Zhang and Tong [14] suggested that aluminum favors easy slip transmission and accommodation among grains. Zhao et al. [15] recently suggested that grain topology (i.e. the geometry of grain boundaries) and micro-texture play an important role in strain heterogeneities. Strain localization behavior in previous experiments [14–16] on specimens with a single layer of grains can be related to the planar variation of grain orientation, since there is no constraint of deformation by other layers of grains. The objective of the present work is to develop a good comprehension of the grain structure and mechanical properties relationship emphasizing the strain localization process in AA5754 alloys. This is done by the use of samples with approximately two layers of grains with sizes in the millimeter range, in order to understand the influence of though-thickness grain interaction on deformation heterogeneity and development of strain localization and to facilitate the comparison between experiments and modeling. 2 Sample preparation and experimental procedure The material under study is a 1 mm thick AA5754 DC sheet with 0.21 wt.% Fe content. The initial grain structure is nearly equiaxed with an average grain size of around 25 μm. Grain structure was revealed under polarizing light after electrical etching with Barker's reagent (48–50% HBF 4 (conc.) 15 ml + H 2 O (distilled) 90 ml). Optical observation was carried out using a Zeiss Optical Microscope. The coarse-grained structure was obtained through a thermomechanical treatment starting with a small rolling reduction pre-deformation (4–10% thickness reduction) followed by annealing at 575 °C for 3 h. The resulting microstructure consists of thin fine-grained layers on the surface and two or more layers of large pancake-shaped grains in the bulk of sheets ( Fig. 1 ). The fine surface grains are removed by grinding. A quasi-2D microstructure was therefore obtained with two layers through the thickness combined with a center-plane separating the two layers when the rolling reduction ranged from 4% to 7%. Samples reduced by 4% had an average post-annealing grain size of about 1.2 mm while samples reduced 7% led to grain sizes of about 0.6 mm. Tensile bars were machined using electron discharge machining (EDM) to ensure high dimensional precision and minimum surface damage. The dimensions of the tensile bars for materials with 1.2 mm and 0.6 mm grain sizes are shown in Fig. 2 (a) and (b) , respectively. A large transition radius is needed to ensure necking within the gauge length. The gauge area reduction was modified compared with the ASTM standard in order to facilitate EBSD and DIC measurements. The initial crystallographic orientation maps on two surfaces of polished tensile bars were obtained using the electron backscattered diffraction (EBSD) technique. EBSD scanning was done at 35× magnification with 15 μm step size over an area slightly larger than the reduced gauge region on both surfaces. EBSD analysis utilizes Kikuchi patterns, which are collected and indexed using an EDAX-TSL OIM system attached to a LEO Scanning Electron Microscope. The display algorithm removes EBSD data which are indexed with a confidence level less than 95% and shows these as black grains. In order to reduce the number of artificial misorientations at the boundaries during large-area scanning, four or eight scans are performed for each surface and a MatLab program was written to merge these scans. The program can detect non-edge and edge grains and determine a mean orientation of non-edge grains and of each pair of corresponding edge grains which cross the boundaries between scans. The experimental error for crystal orientations is within 2° after such treatment under the assumption that the orientation in a grain is homogeneous. This assumption is statistically reasonable since very few little sub-grain structure was detected in hundreds of millimeter sized grains in these samples, after annealing at elevated temperature. Tensile tests were performed at 1 mm/min cross-head speed on a screw-driven tensile machine. Displacement-controlled mode was selected to avoid abrupt changes in deformation rate during necking. Black and white spots were painted on the specimen surface prior to testing. These are used to form a random speckle pattern on the surface. Two high-speed cameras recorded the speckle patterns on both sides of the sample with a time interval of 2 s. Load–displacement curves were recorded and a two-dimensional technique was used to calculate the strain maps on both sides of the tensile bars. A commercial digital image correlation (DIC) system (Aramis) was used to follow the surface deformation pattern during tensile tests. From this we calculate the micro true strain, i.e. normal strain ɛ xx , ɛ yy and shear strain ɛ xy at the facet level. The deformation strain along the thickness direction is estimated based on the assumption of consistent volume during plastic deformation. (This assumption is valid for strains that exceed the elastic strain by a significant margin as is the case here). From these measurements the von Mises equivalent true strain was obtained and used as an index of micro deformation for each facet. A 7 × 5 pixel facet/step size was used for the strain calculation in Aramis, which is equivalent to a real dimension of 125–300/90–200 μm. The accuracy of calculated strain is limited by the stability of facet centers from stage to stage, which is set as 0.04 pixels, leading to a maximum deviation of strain within 0.005. The evolution of grain orientations was measured on one coarse-grained sample during in situ tensile testing at 0.5 mm/min strain rate run in the SEM. In-situ tests were load-controlled. EBSD maps were collected at approximate strains of around 0.06, 0.13 and 0.19. 3 Experimental results 3.1 Samples with an average grain size of 1.2 mm 3.1.1 Grain-level microstructure Grain structures on both sides of the gauge area for Sample A with 1.25 mm grain size are presented in Fig. 3 with inverse pole figure (IPF) color schemes showing the position of tensile axis in the crystal reference frame of grains. In addition, the orientations of individual grains (numbered 1–16) on both sides are represented in the IPF triangle with respect to the tensile axis. 3.1.2 DIC strain mapping The evolution of micro von-Mises equivalent true strain ( ɛ vm ) on both sides at the gauge areas of the sample during deformation is displayed in Fig. 4 (a)–(g) using a color contours (ranging from blue to red), where, the red part of the colorbar defines the range when strains are larger than half peak value ε vm h = ( ε vm p − ε vm m ) / 2 . ε vm p is peak true strain while ε vm m is the medium true strain value. Regions of local strain concentration are colored in red in these maps, and referred to as ‘hot spots’. These nucleate at very early stages of the deformation, and show sustained growth during almost all of the deformation stages. In contrast, the regions with local strain relaxation are colored blue and referred to as ‘cold spots’. These also keep the same positions during almost all deformation stages and deformation stops in these regions after a certain stage. Strain distribution evolutions of a line scan along the tensile direction describes the nucleation and development of ‘hot spots’ from another view. Fig. 4 (h) shows such evolutions along line l in Fig. 3 on both sides parallel to the tensile axis. Strain distribution along this line keeps similar shapes with deformation until the deformation concentrates on one of the initial sharp peaks leading to necking. The evolutions of von Mises true strain distribution on both sides are quite similar, which implies that the grains in Side A have highly coordinated deformation with corresponding grains in Side B, which may also suggest that surface strains calculated from DIC can give a good evaluation of the bulk strains. 3.1.3 Tensile responses of individual grains In order to gain a better understanding of the deformation processes, detailed analyses for individual grains, especially those in the zones of strain concentration and final failure were carried out on Sample A. The initial grain orientations are presented in the form of overlapping Schmid factors on both sides, and a grey-level legend representing the Schmid factors. Strain partitioning in each individual grain was determined from a network of DIC facet points ( Fig. 5 ). Those points of a grain in Side A were selected for analyses so that their corresponding (overlapping in ( x , y ) coordinates) points on Side B are within another grain(s). Therefore, these points are the overlapping points of a pair of grains with each lying on different sides of the sample in order to show if there is coordinated deformation between the corresponding grains. Micro ( ɛ vm )–macro ( ε ¯ vm ) von Mises true strain curves for individual grains are plotted in Fig. 6 . The micro von Mises true strain ɛ vm indicates the strain for each facet point, while macro von Mises true strain ε ¯ vm is the average true strain of all the DIC facet points on one side in the gauge area. The results can be summarized as follows: (a) The deformation inside each grain is quite uniform ( Fig. 6 (a) and (b)) except for the grains which lie at the sample edges, (see Fig. 6 (c), (d) and (f)). The curves for facet points within one edge grain are quite dispersed, indicating that the deformation is quite inhomogeneous. This is due to the different boundary conditions of these grains compared those interior grains which are surrounded and constrained by other grains in the x – y plane. In addition, the ε vm − ε ¯ vm curves of the edge grain are found to have a curvature at different macro strain levels, indicating large variations of the evolution of grain orientations. However, within most of the interior grains, the micro strains maintain a relatively linear relationship with the macro true strains. It seems that grains try to maintain their ‘hardness’ compared to their initial states, indicating that the relatively hardness of (surface) grains is strongly associated with their initial states. (b) As mentioned in Section 3.1.2 , strain maps are quite similar for both sides. Accordingly, the behaviors of the overlap regions corresponding to grains in Side A and B are almost the same ( Fig. 6 (a) and (b)), indicating that the interaction between the pair of grains is necessary to achieve the strain compatibility. (c) The micro von Mises true strains of grains inside the strain localization region ( Fig. 6 (a)–(d)) increases dramatically after necking. Some points of region A8B14 are inside the localization zone which is the starting position for fracture. The ε vm − ε ¯ vm curves show that the local strains of these points increase dramatically after necking, while those of other points in the region stop changing ( Fig. 6 (d)) at some point after necking, implying that deformation inside the same grain may not always be homogenous after necking. (d) Micro von Mises strains of the regions outside the necking zones ( Fig. 6 (e) and (f)) stop increasing after necking. This confirms the achievement of shear localization in such a coarse-grained sample: strain is only concentrated inside the localization region after localized necking. 3.1.4 Surface slip traces Plastic deformation in high stacking fault energy Al alloys is the result of dislocation slip which leaves slip traces observable on polished sample surfaces. Slip trace observations have been made by observing the side surface of the fractured sample as shown in Fig. 7 . In general, one or two slip traces were observed inside each grain, which implies that most of the plastic deformation of grains is contributed by one or two slip systems. Slip traces with relatively homogeneous interspacing were observed within the body of the grains ( Fig. 7 (c)). Transition areas with mixed slip traces were recorded near grain boundaries. These observations suggest that while plastic deformation is relatively homogeneous in the grain interior, the constraints imposed by grain to grain compatibility demands the involvement of additional slip systems near the grain boundaries. The slip traces are the lines of intersection between the slip planes of activated slip systems and the sample surface. According to the Schmid law, deformation occurs on the slip system with the largest Schmid factor. This is especially true in the case where there is only one set of slip traces parallel to each other on the surface of a grain. The observed slip traces majorly agree with the predicted slip systems. 3.1.5 The evolution of grain orientations The initial surface grain structure and the evolution of grain orientations during an in situ tensile test on Sample B are represented using IPF color maps ( Fig. 8 (a) and (b) ) and IPFs with respect to the tensile axis ( Fig. 8 (c)–(f)). The distribution of grain orientations becomes more random with increasing strain. The division of grain orientation inside one grain is observed in parts with different neighboring oriented grains. In general, the lattice rotates to achieve deformation axis closer to 〈0 0 1〉, 〈1 1 2〉 or 〈1 1 1〉. 3.2 Sample with an average grain size of 0.6 mm 3.2.1 Grain-level microstructure To further understand grain-grain interaction during deformation, a tensile test was performed on Sample C with shape II, which provides more grains in the gauge area compared to Sample A(B) with larger grain sizes. Grain structure for both sides of the gauge region for Sample C with average grain size of 0.6 mm is presented in IPF color maps with respect to tensile axes ( Fig. 9 ). 3.2.2 DIC strain mapping The observations in Sample C were similar to those in Sample A ( Fig. 10 ): a few ‘hot spots’ form at the beginning, and grow in a sustained manner until facture occurs at one of them. In addition, ‘cold spots’ were also observed from the beginning and these eventually see a stop in increasing strain after a certain stage. The half peak value of the micro von Mises true strain is almost the same as that in Sample A for a specific macroscopic strain level, which may due to using the same facet size compared to grain size to calculate strain. 3.2.3 Tensile responses of oriented grain clusters in Sample C The local tensile responses of grain clusters (including a few surface grains) were investigated on Sample C. As shown in Fig. 11 , grain orientations were indexed by the overlapping IPF coloring maps of both sides. According to the spot size of ‘hot and cold spots’, the cluster of grains has a diameter around 700 μm, including 2–4 grains in each side. Strain partitioning of each oriented grain cluster was determined using a network of DIC facet points. Fig. 12 (a) and (b) show the ‘hot spot’ regions, (c) and (f) are within the ‘cold spots’ and (d) and (e) have comparable high strain with their surroundings as marked in Fig. 11 . The initial grain-level microstructure is represented by the overlapping IPF color maps on both sides. The color of grains refers to the position of the tensile axis in the crystal reference frame. It is interesting that strain concentration regions are often associated with grain clusters including green grains (i.e. grains of ∼〈1 0 1〉 direction close to tensile axis) in both sides while the purple grains (i.e. grains of ∼〈1 1 2〉 direction close to tensile axis) always leads to the strain relaxation regions. 3.2.4 Grain orientations of fractured Sample C EBSD maps were obtained on both sides of the fractured sample for Sample C. ( Fig. 13 ) In general terms, grains tend to be in purple and blue within the IPF maps, representing orientations that concentrate near the 〈0 0 1〉–〈1 1 1〉 line after tensile deformation, which could be predicted from the evolution rule observed in Sample B. 4 Discussion Although grain behavior is not solely controlled by crystal orientation, it seems that grains of ∼〈1 0 1〉 direction, e.g. the green grains such as grain No. 11, 14 and 15 on Side A as shown in Fig. 3 , have more possibility to become the region of ‘hot spots’ with large local strains, while the grains of ∼〈1 1 2〉 directions, i.e. the purple grains in Fig. 3 (No. 5 in Side A and No. 4, 6, and 9 in Side B), usually see small strains and become the regions of ‘cold spots’. Furthermore, oriented grain clusters in Sample C that include green grains are the preferred site for a local strain concentration region, while grain clusters including purple grains always show-up as strain relaxation regions. As shown in Fig. 14 , the initial local deformation tends to be larger with decreasing misorientation between tensile axis and the 〈1 0 1〉 axis of a grain. From the above observations, it appears that the grains of ∼〈1 0 1〉 direction are softer compared to all other orientations, while the grains of ∼〈1 1 2〉 direction are harder to deform. The relative strengths of grains during deformation are often correlated with their Taylor factors [16] or Schmid factors [14,17] . A grain is relatively soft with a smaller Taylor factor or larger Schmid factor. Actually the Schmid factor is almost the inversion of Taylor factors for the grains with orientations of high symmetry including 〈0 0 1〉 and 〈1 1 1〉, while there is big variation between the Taylor factor and the inverse of the Schmid factor for grains with ∼〈1 0 1〉 direction (see Fig. 15 ). These two high symmetry orientations are quite stable during the deformation of high stacking fault energy fcc materials [18] . The construction of the yield surface of a textured polycrystalline material is based on the average Taylor factors of a representative cluster of grains which are deformed in different deformation modes [19] . The deformation heterogeneity in a plane strain compression test of a large grained sample has been assessed based on Taylor factors by Raabe et al. [16] , while it has been presented by Zhang and Tong [14] and Tatschl and Kolednik [17] based on Schmid factor in the uniaxial tension test of large grained samples. From the Schmid and Taylor factor contours ( Fig. 15 ) that are plotted on the tensile direction in the IPF triangle, it shows that the grains with both 〈1 0 1〉 and 〈1 1 2〉 directions have relatively high Schmid factors of around 0.42. The Taylor factors of grains with 〈1 0 1〉 orientation (∼3.7), however, is higher than that with 〈1 1 2〉 orientations (∼3), indicating the grains with 〈1 0 1〉 should be harder. The deformation of a grain is also constrained by the surrounding grains and the corresponding grain on the other side. For example, there is quite a large variation of deformations for the grains of similar orientations, e.g. ∼〈1 0 1〉 (see Fig. 14 ), due to this effect. We have shown that there is coordinated deformation between the grain of one layer and the corresponding grain on another layer. It could be that how much the two corresponding grains can deform depends on the overall orientation relationship of these two grains. To illustrate this point, average cross-sectional (TD × ND) Schmid and Taylor factors are plotted ( Fig. 16 ) in reference to the x 1 coordinates along the loading direction (RD), where the Schmid and Taylor factors are averaged over all the points of each TD × ND section along the loading direction. The approximate position of localized necking and final failure is also noted in the curves. It can be seen that the position of localized necking and final failure correlates well with the position corresponding to maximum Schmid factor. In addition, for Sample C, the regions with maximum Schmid factors are located in the same regions with minimum Taylor factors due to statistical effect of more grains. From the above discussion, it seems that the relative ‘hardness’ of grains in uni-axial tension is less correlated with the initial Taylor factors than with Schmid factor as shown in earlier works [14,17] . This may be due to the additional freedom for surface deformation provided by these extra-large grains, leading to only the one or two slip systems that have the largest Schmid factors being activated at the surface. The strain compatibility between the surface grains can be accomplished by the activation of more slip systems at the regions near the grain boundaries. In a typical material that has many grains through the thickness, the internal grains in the bulk are fully constrained by neighbors. More slip systems need to be activated to provide strain compatibility between neighboring grains, in this case the relative ‘hardness’ of the grains may be more strongly coupled with their Taylor factors. It should also be noted that this discussion is based on the initial orientation of grains. During plastic deformation, the grains can re-orient which may change the relative strength of grains leading to geometrical softening and strain localization [20] . Early research [21] on the statistical distribution of grain orientations suggested that texture components see no significant changes before strain localization. However, our analysis based on individual grains demonstrated the preferred evolution trend: grains with 〈1 0 1〉 parallel to the tensile axis will generally rotate toward the 〈0 0 1〉–〈1 1 1〉 line in the IPFs with respect to the tensile axis. Similar observations are obtained from X-ray diffraction results of tensile tested sample [22,23] and SEM in situ tensile testing [17] of polished samples with a single layer of large grains. Furthermore, due to the effect of deformation heterogeneity and anisotropy, the local deformation mode can deviate from the ideal applied uniaxial tension deformation model, e.g. Δ ε = 1 0 0 0 − 0.5 0 0 0 − 0.5 Δ ε 11 which is usually used to calculate the initial Taylor factor for uniaxial tension. Thus, the relative ‘hardness’ of grains can change during deformation. For example, the Taylor factors in the grains with 〈1 0 1〉 close to tensile deformation can become smaller during straining. Although the crystal re-orientation can be measured by in situ EBSD measurement during tension and the deformation mode on the surface can be tracked by DIC, applying both technologies at the same time to a reasonable resolution and precision is still too challenging. These effects, however, can be considered by finite element modeling, which will be discussed in a separate paper where either a Taylor factor-based formulation, i.e. the FE-TBH model [24] , or the conventional viscous crystal plasticity formations (CPFEM) are used to represent the behaviors of crystalline grains. The interaction between grains is quite hard to be analyzed experimentally in a quantitative manner. However, microstructure-based models could trace both orientation and strain evolution of grains and pick up the grain interaction effect. The simulation results of the nucleation and growth of ‘hot spots’ by both the FE-TBH and CPFEM models show reasonable agreement with experimental results and support some new micro-features of strain distribution on the sub-grain level. Fig. 17 shows examples of deformation of individual grains predicted by the CPFEM calculation [25] . The strain distribution inside one grain is relatively even with fluctuations at grain boundaries due to the interaction with neighbors for both layers. 5 Conclusions The micro tensile responses of coarse-grained AA5754 tensile bars have been reported in this paper. Strain distribution evolution maps present the special deformation processes of coarse-grained samples: ‘hot bands’ nucleate very early during deformation and most of them keep growing throughout deformation stages. Two-surface von Mises true strain maps reveal that grains belonging to different layers deform in tandem through the sample thickness. Micro–macro von Mises true strain curves provide evidence for strain localization: strain localization regions can be attributed to dramatically increasing strain after a certain stage while deformation stops in all other regions of the sample. Linking the observed deformation behavior with microstructure, the following conclusions can be drawn: (a) the average initial cross-sectional microstructure (Taylor or Schmid factor) factor can be used to predict several possible sites for strain concentration; (b) ‘hot spots’ occur in local regions with soft grains; (c), those grains with 〈1 0 1〉 close to the tensile direction are favored sites for strain concentration; (d), grains rotate toward the 〈0 0 1〉–〈1 1 1〉 line in IPF triangle of the tensile axis and change the relative ‘hardness’ of grains, which is probably the determining factor for the sustained growth of ‘hot spots’. Acknowledgements We are very grateful to Mr. Robert Kubic, General Motors R&D Center, for helping with the EBSD work. The financial support from General Motors of Canada and the Natural Sciences and Engineering Research Council (NSERC) of Canada is acknowledged. References [1] J. Hirsh Materials Forum 28 2004 15 23 [2] The Canadian Aluminium Industry Technology Roadmap, Industry Canada, report, 2000, edition. [3] J Sarkar T.R.G. Kutty D.S. Wilkinson J.D. Embury D.J. Lloyd Aluminium Alloys: Their Physical and Mechanical Properties 331–333 Pts 1–3 2000 583 588 [4] J. Kang M. Jain D.S. Wilkinson D. Embury S. Kim A.K. Sachdev Tenth International Symposium on Plasticity and Its Current Application 2003 NEAT Press pp. 181–183 [5] L. Anand W.A. Spitzig Journal of the Mechanics and Physics of Solids 28 1980 113 128 [6] A. Korbel P. Martin Acta Metallurgica 36 9 1988 2575 2586 [7] J.C. Huang G.T. Gray III Acta Metallurgica 37 12 1989 3335 3347 [8] W.B. Lee K.C. Chan Textures and Microstructure 14–18 1991 1221 1226 [9] P.D. Wu A. Graf S.R. MacEwen D.J. Lloyd M. Jain K.W. Neale Internaional Journal of Solids and Structures 42 2005 2225 2241 [10] X.H. Hu M. Jain D.S. Wilkinson R.K. Mishra Acta Materialia 56 13 2008 3187 3201 [11] H.-S. Zhang K. Komvopoulos Philosophical Magazine 90 16 2010 2235 2248 [12] Braunschweig, Aramis V4.7 Manual. 2001: GOM mbH. [13] D. Raabe M. Sachtleber Z. Zhao F. Roters S. Zaefferer Acta Materialia 49 17 2001 3433 3441 [14] N. Zhang W. Tong International Journal of Plasticity 20 2004 523 542 [15] Z. Zhao M. Ramesh D. Raabe A. Cuitino R. Radovitzky International Journal of Plasticity 2008 [16] D. Raabe M. Sachtleber H. Weiland G. Scheele Z. Zhao Acta Materialia 51 6 2003 1539 1560 [17] A. Tatschl O. Kolednik Materials Science and Engineering A 342 2003 152 168 [18] W. Mao Y. Yu Materials Science and Engineering A 367 2004 277 281 [19] P.V. Houtte S. Li M. Seefeldt L. Delannay International Journal of Plasticity 21 2005 589 624 [20] I.L. Dillamore J.G. Roberts A.C. Bush Metal Science 13 2 1979 73 77 [21] D. Peirce R.J. Asaro A. Needleman Acta Metallurgica 30 1982 1087 1119 [22] L. Margulies G. Winther Science 291 2001 2392 2394 [23] G. Winther L. Margulies H.F. Poulsen S. Schmidt A.W. Larsen E.M. Lauridsen S.F. Nielsen A. Terry Materials Science Forum 408–412 2002 287 292 [24] X.H. Hu G.A. Cingara D.S. Wilkinson M. Jain P.D. Wu R.K. Mishra Materials and Continua 14 2009 97 122 [25] X.H. Hu, G.-Z. Zhu, D.S. Wilkinson, M. Jain, P.D. Wu and R.K. Mishra, Finite element and experimental studies of tensile deformations of AA5754 sheet alloys with extra-large grains, unpublished.
What problem does this paper attempt to address?
-
Strain Ratio Effects on Low-Cycle Fatigue Behavior and Deformation Microstructure of 2124-T851 Aluminum Alloy
Hong Hao,Duyi Ye,Chuanyong Chen
DOI: https://doi.org/10.1016/j.msea.2014.03.040
2014-01-01
Abstract:The low-cycle fatigue tests of 2124-T851 aluminum alloy with strain ratios of −1, −0.06, 0.06 and 0.5 were conducted under constant amplitude at room temperature. Microstructural and fractographic examinations of the material after fatigue tests were performed by optical microscopy (OM) and scanning electron microscopy (SEM), respectively. Firstly, the results showed that the material exhibited cyclic softening characteristic as a whole. The degree of softening decreased linearly with the increasing strain amplitude and the decreasing strain ratio. The lower fatigue life and ductility of the material corresponded to the larger strain ratios. Secondly, microstructure observations revealed that the density and length of slip bands increased with the increasing strain ratio at the given strain amplitude, and so did the volume fraction and size of coarse constituents, which were responsible for the reduction of fatigue life and ductility of the material. Finally, the SEM micrographs revealed that multiple crack initiation sites took place on the fracture surfaces at different strain ratios. The reduction of stable crack growth area with the increasing strain ratio was observed. Unstable crack growth region was only observed under R≠−1.
-
Strain inhomogeneity and grain refinement in AA6063 aluminum alloy subjected to repetitive corrugation and straightening: comparison of experimental and finite element analysis
N. Thangapandian,J. Jenix Rino,Balasivanandha Prabu Shanmugavel
DOI: https://doi.org/10.1007/s12289-022-01649-9
2022-02-07
International Journal of Material Forming
Abstract:The strain distributions in aluminum alloy AA6063 resulting from Repetitive Corrugation and Straightening (RCS) at room temperature were examined experimentally and also through finite element simulation. The hardness and tensile strength of the material after every passes are reported. The local strain in the samples were determined using circle grid analysis method. EBSD images confirmed the presence of sub-grains inside the coarse grains. Hardness measurements showed non-uniform values along the length of the specimen, revealed the presence of non-uniform strain as well as grain refinement. After 8 passes, the hardness was above twice that of the starting materiali, i.e. (49 Hv in the starting material to 98Hv after 8 passes). The tensile strength increased up to 6 passes from 115 MPa to 204 MPa, and decreased beyond that due to the formation of surface cracks. Finite element analysis (FEA) was carried out under identical experimental conditions using ABAQUS 6.10 to estimate the stress strain distributions up to 10 passes. The RCS die of width ~ 3 mm and groove angle 30˚ and the flat die was designed as the discrete rigid, since the deformation on the die is of no interst. The FEA reveals that the strain increases after each passes displays consistency with the experimental results, and a lack of homogeneity in deformation.
materials science, multidisciplinary,metallurgy & metallurgical engineering,engineering, manufacturing
-
Microstructural and texture evolution investigation of laser melting deposited TA15 alloy at 500?C using in-situ EBSD tensile test
Muhammad Rizwan,Junxia Lu,Rafi Ullah,Yuefei Zhang,Ze Zhang
DOI: https://doi.org/10.1016/j.msea.2022.144062
2022-01-01
Abstract:The deformation behavior in polycrystalline Laser melting deposited (LMD) TA15 alloy was investigated at a temperature of 500 degrees C using advanced in-situ Secondary electron microscopy (SEM) tensile setup coupled with Electron backscatter diffraction (EBSD). The grains morphology, crystal orientation, preferred pole texture, activated slip traces, grain boundaries evolution, and strain-induced misorientation were systematically analyzed during tensile straining. It was determined that the diverse alpha-grains (crystal orientation and morphology) revealed inhomogeneous deformation under tensile loading. The overall strain was allocated by grain rotation and substructure formation. The preferred oriented (0001) pole texture maximum was increased during initial strain level. In contrast, at higher strain, the texture intensity was decreased due to the formation and relocation of the newly formed substructures. The fraction of low angle grain boundaries (LAGBs 2-10 degrees) was increased continually with increasing strain. Strain-induced dislocations within larger grains formed new LAGBs, and the increasing fraction of LAGBs was validated by Kernel average misorientation (KAM) map. The prismatic slip mode was identified as the primary active slip system at the initial deformation stage. The fraction of basal and pyramidal < a > slips increased as the strain increased, describing the opposite trend to prismatic slip. The grains rotation and strain accumulation at the grain boundaries accommodated the plastic deformation, ultimately causing boundary sliding.
-
In Fl Uence of Texture and Grain Re Fi Nement on the Mechanical Behavior of AA 2219 Fabricated by High Shear Solid State Material Deposition
O. G. Rivera,P. G. Allison,L. N. Brewer,O. L. Rodriguez,J. B. Jordon,T. Liu,W. R. Whittington,R. L. Martens,Z. Mcclelland,C. J. T. Mason,L. Garcia,J. Q. Su,N. Hardwick
DOI: https://doi.org/10.1016/j.msea.2018.03.088
IF: 6.044
2018-01-01
Materials Science and Engineering A
Abstract:Issues with rapid grain growth, hot cracking and poor ductility have hindered the additive manufacturing and repair of aluminum alloys. Therefore, this is the first investigation to spatially correlate the processing-structureproperty relations of a precipitation hardened aluminum alloy 2219 (AA2219) material with respect to deposition orientations and build layers. The AA2219 material was processed by a high deposition rate (1000 cm(3)/ h) solid-state additive deposition process known as Additive Friction Stir Deposition or MELD. An equiaxed grain morphology was observed in the three orientations, where Electron Backscatter Diffraction (EBSD) identified a layer-dependent texture with a strong torsional fiber A texture in the top of the build transitioning to weaker textures in the middle and bottom layers. Interestingly, the tensile behavior reflected the texture layer-dependence with tensile strength increasing from the bottom to the top of the deposition. However, there were no statistically significant differences in hardness measured from the top to the bottom of the deposition. Furthermore, no orientation dependence on mechanical properties was observed for compression and tension specimens tested at quasi-static (0.001/s) and high (1500/s) strain rate. Transmission Electron Microscopy (TEM) determined a lack of theta(,) precipitates in the as-deposited cross-section, therefore resulting in no precipitation strengthening.
-
Deformation-Induced Surface Roughening of an Aluminum–Magnesium Alloy: Experimental Characterization and Crystal Plasticity Modeling
Yannis P. Korkolis,Paul Knysh,Kanta Sasaki,Tsuyoshi Furushima,Marko Knezevic
DOI: https://doi.org/10.3390/ma16165601
IF: 3.4
2023-08-13
Materials
Abstract:The deformation-induced surface roughening of an Al-Mg alloy is analyzed using a combination of experiments and modeling. A mesoscale oligocrystal of AA5052-O, obtained by recrystallization annealing and subsequent thickness reduction by machining, that contains approx. 40 grains is subjected to uniaxial tension. The specimen contains one layer of grains through the thickness. A laser confocal microscope is used to measure the surface topography of the deformed specimen. A finite element model with realistic (non-columnar) shapes of the grains based on a pair of Electron Back-Scatter Diffraction (EBSD) scans of a given specimen is constructed using a custom-developed shape interpolation procedure. A Crystal Plasticity Finite Element (CPFE) framework is then applied to the voxel model of the tension test of the oligocrystal. The unknown material parameters are determined inversely using an efficient, custom-built optimizer. Predictions of the deformed shape of the specimen, surface topography, evolution of the average roughness with straining and texture evolution are compared to experiments. The model reproduces the averaged features of the problem, while missing some local details. As an additional verification of the CPFE model, the statistics of surface roughening are analyzed by simulating uniaxial tension of an AA5052-O polycrystal and comparing it to experiments. The averaged predictions are found to be in good agreement with the experimentally observed trends. Finally, using the same polycrystalline specimen, texture–morphology relations are discovered, using a symbolic Monte Carlo approach. Simple relations between the Schmid factor and roughness can be inferred purely from the experiments. Novelties of this work include: realistic 3D shapes of the grains; efficient and accurate identification of material parameters instead of manual tuning; a fully analytical Jacobian for the crystal plasticity model with quadratic convergence; novel texture–morphology relations for polycrystal.
materials science, multidisciplinary,chemistry, physical,physics, applied, condensed matter,metallurgy & metallurgical engineering
-
Strain Distribution and Lattice Rotations During In-Situ Tension of Aluminum with a Transmodal Grain Structure
W. Q. Gao,C. L. Zhang,M. X. Yang,S. Q. Zhang,D. Juul Jensen,A. Godfrey
DOI: https://doi.org/10.1016/j.msea.2021.142010
IF: 6.044
2021-01-01
Materials Science and Engineering A
Abstract:Samples of aluminum were prepared using spark plasma sintering from a mixture of coarse (average particle size of 6 mu m) and fine (average particle size of 1 mu m) powders to achieve a heterogeneous transmodal grain size distribution covering a range of grain sizes from approximate to 1 to 10 mu m. By careful choice of surface markers both electron back-scatter diffraction (EBSD) data, to track crystal rotations, and digital image correlation (DIC) data, to track local plastic deformation, were collected from the same region during in-situ tensile deformation up to a strain of epsilon t = 0.126. A heterogeneous pattern of crystal rotation is observed for all grain sizes, although in some smaller grains (defined as those <4 mu m diameter) no clear grain sub-division was identified. Plastic strain was more concentrated in the larger grains, but the average rotation rate of the smaller grains was found to be higher than that of larger grains, showing also a much wider spread in rotation rate. Based on the change in average orientation, a clear orientation dependence in the tensile axis rotation direction was observed for many larger grains, in agreement with previously reported data for aluminum with average grain size of 75 mu m, whereas the smaller grains showed a more complex rotation behavior, with more of these grains showing unexpected tensile axis rotations. The combination of both EBSD and DIC during in-situ experiments provides a rich data set for analysis of plastic deformation in samples with a heterogeneous microstructure.
-
Effect of second-phase precipitates on deformation microstructure in AA2024 (Al–Cu–Mg): dislocation substructures and stored energy
Daniel Irmer,Can Yildirim,Mohamed Sennour,Vladimir A. Esin,Charbel Moussa
DOI: https://doi.org/10.1007/s10853-024-10205-6
IF: 4.5
2024-10-21
Journal of Materials Science
Abstract:The importance of comprehensive multiscale characterisation in advancing our understanding of engineering materials is undeniable but remains a challenging pursuit. Combining complimentary microstructure characterisation techniques, including transmission electron microscopy, electron backscatter diffraction and dark-field X-ray microscopy (DFXM), the formation of deformation microstructures is investigated in presence of shearable and non-shearable hardening precipitates in an industrial aluminium alloy (AA) 2024 (Al–Cu–Mg family). The alloy was used in naturally aged T3 (with shearable co-clusters and Guinier–Preston–Bagaryatsky (GPB) zones) and peak-hardened T8 (with non-shearable S-phase precipitates) states. After cold rolling with thickness reductions varying from 25 to 60% (or corresponding von Mises strain from 0.33 to 1.06), the T8 state revealed a higher sub-boundary density with slightly smaller mean disorientation angle, as compared to those in the T3 state. At a von Mises strain of 0.33, the T8 state exhibited higher long-range orientation gradients, as compared to the T3 state, for higher strain orientation gradients in T3 surpass those in T8 state. With DFXM, distinct 3D substructures are shown, revealing ellipsoidal sub-grains in the T8 state and pancake-like sub-grains in the T3 state. Moreover, the stored energy induced by cold rolling is higher for the T8 state. These results indicate different deformation microstructures, formed in the same AA2024 but hardened by shearable and non-shearable precipitates.
materials science, multidisciplinary
-
Anisotropic Mechanical Behavior and Microstructural Characteristics of AA2099-T83 Alloy Sheet: Experiment and Simulation
Mengdi Li,Lingguo Zeng,Weijiu Huang,Xusheng Yang,Li Hu,Daiyu Xiong,Xianghui Zhu,Xin Wang
DOI: https://doi.org/10.1007/s11665-024-09738-3
IF: 2.3
2024-06-26
Journal of Materials Engineering and Performance
Abstract:Mechanical anisotropy remains a highly regarded topic in the application of Al-Li alloy sheet. The microcosmic factors affecting mechanical anisotropy and the deformation characteristics within and between grains of this alloy during tensile deformation still require further elucidation. In this study, uniaxial tensile experiments were conducted on AA2099-T83 Al-Li alloy sheet along different directions, and the mechanical response, as well as the variations in stress and strain within and between grains, was predicted using the crystal plasticity finite element method (CPFEM). The results show that AA2099-T83 sheet display significant anisotropy in yield strength and elongation. The uneven distribution of the T 1 phase predominantly influences strength anisotropy, whereas texture and grain morphology have a lesser impact. CPFEM simulations at a true strain of 0.06 reveal that texture significantly affects elongation anisotropy, with the strain distribution of grains with a preferred orientation in 0° and 90° tensile directions aligning with the model's overall strain distribution pattern. Notably, samples stretched at 90° demonstrate a higher strain gradient between grains, potentially increasing the likelihood of crack initiation.
materials science, multidisciplinary
-
Study of the recrystallization behaviour of the aluminium 1565ch alloy during hot rolling of the as cast structures
Evgenii Aryshenskii,Juergen Hirsch,Vasiliy Yashin,Sergey Konovalov,Ekaterina Chitnaeva
DOI: https://doi.org/10.1088/2053-1591/ab13b6
IF: 2.025
2019-04-10
Materials Research Express
Abstract:The recrystallization process in 1565 ch aluminium alloy samples was studied in this paper. Existing and well-studied AA5182 aluminium alloy was also investigated for comparison. Samples of both alloys were cut from pre-homogenized ingots and rolled in different modes. After rolling the samples were annealed at different temperatures and the resulting microstructure was studied using scanning electron microscopy and optical microscopy, as well as EBSD (Electron backscatter diffraction) analysis. The study revealed significant differences in the recrystallization processes of these two alloys. Signs of the recrystallization onset without grain growth process can be observed at 10% deformation in 1565 ch alloy. At the same time signs of recrystallization in AA5182 can been observed only when the degree of deformation is over 30%, but the process itself is much faster. 5182 alloy fully recrystallizes after all hot rolling schedules, except for those with 10% deformation. In such case recrystallization timing as a function of Zener-Hollomon parameter and prior strain level ranges from a few seconds to half an hour. In major cases 1565 ch alloy has partially recrystallized structure, which takes 10 to 95 volume %. Exception is the schedule with deformation 50% and Zener parameter 1.77 × 1015, where full recrystallization occurs. With such strain parameters 1565 alloy grain size reaches 6–14 μm. At the same time similar processing provides larger grain size in 5182 alloy, it reaches 10–20 μm. Substructure review immediately after deformation demonstrated, that cellular polygonised structure forms faster in 1565 ch alloy, which can be explained by higher magnesium content. In both cases PSN nucleation mechanism prevails. However, other grain nucleation mechanisms can be observed in 5182 alloy. The main reasons for the abovementioned phenomenon are different magnesium contents and volumes of fine secondary phases.
materials science, multidisciplinary
-
Evolutions of Grain Orientation and Dislocation Boundary in Aa1050 Aluminum Alloy During Cold Rolling from Low to Medium Strains
Liu Qing,Yao Zongyong,A. Godfrey,Liu Wei
DOI: https://doi.org/10.3321/j.issn:0412-1961.2009.06.001
IF: 1.797
2009-01-01
ACTA METALLURGICA SINICA
Abstract:The dislocation structure evolution in AA1050 aluminum alloy during cold rolling from low to medium strains was investigated using electron channeling contrast, (ECC) imaging and the electron backscattered diffraction (EBSD) techniques. The results show that the grains are subdivided into a typical cell-block structure and there is a strong correlation between deformation microstructure and grain orientation. Based on the characterizations of grain subdivision and dislocation boundary structure, grains can be classified into three types: Type A-grains Containing two sets of geometrically necessary boundaries (GNBs), Type B-grains containing one set, of GNBs, and Type C-grains consisting of large dislocation cells structure. Most of grains with Copper, Brass and Goss orientations have Type A microstructure; grains with S orientation have Type B microstructure grains with Cube orientation have Type C microstructure. The alignment of, the extended dislocation boundaries depends strongly on the grain orientation. In most grains the boundaries have inclination angles of +/-(30 degrees-40 degrees) to rolling direction (RD), and are approximately parallel to the trace's of the most active {111} slip planes as identified by a Schmid factor analysis.
-
Effect of Initial Orientation on the Anisotropy in Microstructure and Mechanical Properties of 2195 Al-Li Alloy Sheet during Hot Tensile Deformation
Jian Ning,Jiangkai Liang,Xinyu Hu,Xianggang Ruan,Zhubin He
DOI: https://doi.org/10.3390/ma16145012
2023-07-15
Abstract:The 2195 Al-Li alloy, as one of the representative third-generation Al-Li alloys, has extensive applications in lightweight aerospace structures. In this paper, the anisotropy in mechanical properties and microstructure evolution of 2195 Al-Li alloy sheets were investigated under a strain rate of 0.01, 0.1, 1 s-1 and a temperature of 440 and 500 °C. Experimental results showed that the hot tensile properties of the 2195 Al-Li alloy sheet exhibited a strong dependence on loading directions. The peak stress (PS) and elongation (EL) along the rolling direction (RD) were larger than the transverse direction (TD). For the tests carried out at 440 °C-1 s-1, the PS values of the sheets stretched along the RD and TD are 142.9 MPa and 110.2 MPa, respectively. And, most of the PS anisotropy values are larger than 15%. The anisotropy in EL is less significant than in PS. All the differences are about 10%. Moreover, dimples in the samples stretched along RD were more and deeper than those along TD at 440 °C. The fracture morphology along RD and TD were similar, and both were cleavage fractures at 500 °C. Particularly, the fractions of high angle grain boundaries (HAGBs) along TD were all about 5% larger than those of RD. And, there were more small-sized continuous dynamic recrystallization (CDRX) grains inside the initial grains and discontinuous dynamic recrystallization (DDRX) grains featured with the local bulge of grain boundaries along TD. This was due to the smaller average Schmid factor and the vertical EL trend of the initial grains when the samples were stretched along TD. A model of grain evolution during the dynamic recrystallization (DRX) along RD and TD was proposed based on EBSD results. The Schmid factor and banded structure had a more prominent effect on the hot ductility of the 2195 Al-Li alloy compared with the degree of DRX, thus presenting a higher EL and better hot ductility along RD.
-
Subgrain Size Modeling and Substructure Evolution in an AA1050 Aluminum Alloy during High-Temperature Compression
Qi Yang,Tomasz Wojcik,Ernst Kozeschnik
DOI: https://doi.org/10.3390/ma17174385
IF: 3.4
2024-09-05
Materials
Abstract:For materials with high stacking fault energy (SFE), such as aluminum alloys, dynamic recovery (DRV) and dynamic recrystallization (DRX) are essential softening mechanisms during plastic deformation, which lead to the continuous generation and refinement of newborn subgrains (2° ˂ misorientation angle ˂ 15°). The present work investigates the influence of compression parameters on the evolution of the substructures for a 1050 aluminum alloy at elevated temperatures. The alloy microstructure was investigated under deformation temperatures ranging from 300 °C to 500 °C and strain rates from 0.01 to 0.1 s−1, respectively. A well-defined substructure and subsequent subgrain refinement provided indication of the evolution laws of the substructure under high-temperature compression. Corresponding experimental data on the average subgrain size under various compression conditions were obtained. Two different independent average subgrain size evolution models (empirical and substructure-based) were used and applied with several internal state variables. The substructure model employed physical variables to simulate subgrain refinement and thermal coarsening during deformation, incorporating a corresponding dislocation density evolution model. The correlation coefficient (R) and root mean square error (RMSE) of the substructure-based model were calculated to be 0.98 and 5.7%, respectively. These models can provide good estimates of the average subgrain size, with both predictions and experiments reproducing the expected subgrain size evolution using physically meaningful variables during continuous deformation.
materials science, multidisciplinary,chemistry, physical,physics, applied, condensed matter,metallurgy & metallurgical engineering
-
Multi-scale modeling of decohesion characteristics of second phase particles from the matrix in uniaxial tension in a high strength aluminum alloy
Abhishek Sarmah,Mukesh K. Jain
DOI: https://doi.org/10.1016/j.engfracmech.2024.110013
IF: 5.4
2024-03-12
Engineering Fracture Mechanics
Abstract:This research investigates stress evolution and plastic deformation characteristics influencing particle–matrix interface decohesion in AA7075-O aluminum sheets. Employing a combined molecular dynamics (MD) - finite element (FE) approach, three interfaces (Al-η, Al-θ, Al-Fe-rich intermetallic) are studied under various loading conditions. Cohesive properties, represented by traction-separation (T-S) curves, are derived using a novel methodology from MD simulations to estimate critical peak traction ( t ) initiating decohesion and the work of separation ( G ) at quasistatic strain rates. It is shown that all three interfaces are weaker in shear than in normal loading. The cohesive property estimates from MD simulations are utilized as input in FE-based models with both real and simplified microstructures of AA7075-O sheet material. These models are subjected to large plastic strains, and the decohesion behavior of each particle type are analyzed. It is shown that particle decohesion is a function of inherent cohesive properties, local inter-particle alignment with respect to loading direction, particle morphology and particle size. Interparticle alignment between 0 and 45 degree promote particle decohesion. Larger sized particles with smaller dimensions along loading direction aid early decohesion of particles. Decohesion of a particle can also facilitate debonding of neighbouring particles under continued loading. Fe-rich particles have higher likelihood for decohesion due to their weaker interface. The θ precipitates, despite having comparable interface strength as η particles, manifest a tenfold increase in susceptibility to decohesion due to the effect of particle morphology and size.
mechanics
-
A Study on Peripheral Grain Structure Evolution of an AA7050 Aluminum Alloy with a Laboratory-Scale Extrusion Setup
Yiwei Sun,Xiaolong Bai,Daniel Klenosky,Kevin Trumble,David Johnson
DOI: https://doi.org/10.1007/s11665-019-04208-7
IF: 2.3
2019-07-17
Journal of Materials Engineering and Performance
Abstract:A laboratory-scale hot extrusion setup was designed to investigate recrystallization and grain growth behavior of an AA7050 alloy during extrusion and subsequent heat treatments. Compared with industrial extrusion, the laboratory-scale process enabled rapid water quenching of extrudate with less delay so that the dynamic grain structure development was captured. After extrusion, static microstructure evolution in the extrudates was studied using salt bath annealing for 5 and 15 s at 490 °C and solutionization treatment for 1 h at 490 °C. The salt bath annealing was a simulation of the delay of press quenching in typical industrial extrusion practices. In the as-quenched extrudates, the peripheral region mainly exhibited continuous dynamic recrystallization and geometric dynamic recrystallization, whereas in the core region discontinuous dynamic recrystallization dominated. A <100> and <111> double fiber texture was identified in extrudates, and recrystallization behavior was found to be orientation dependent. The <100> oriented grains contained more sub-grain boundaries and better-defined sub-grains and had a higher tendency to fragment via continuous recrystallization, while the <111> oriented grains produced less sub-grain boundaries and did not recrystallize. Subsequent heat treatments resulted in static recrystallization and abnormal growth of the continuously recrystallized grains. Additionally, the effects of extrusion temperature (440, 480 and 520 °C) and punch speed (0.7, 1.4 and 2.1 mm/s) on grain structure were discussed. A revised grain structure evolution mechanism based on the observation of 7050 extrusion was proposed.
materials science, multidisciplinary
-
Microstructure and Texture Evolutions of Aa1050 Aluminum Alloy Cold Rolled to High Strains
Yao Zongyong,Liu Qing,A. Godfrey,Liu Wang
DOI: https://doi.org/10.3321/j.issn:0412-1961.2009.06.002
IF: 1.797
2009-01-01
ACTA METALLURGICA SINICA
Abstract:The evolutions of the microstructure and texture of AA1050 alloy cold rolled to large strains have been investigated using electron channeling contrast (ECC) imaging and electron backscattered diffraction (EBSD) techniques. It is found that the microstructure evolves from a cell block structure at low strains into a lamellar structure at, high strains, within which most of lamellar boundaries (LBs) are parallel to rolling direction (RD). Two mechanisms contribute to the microstructure transition, i.e., a gradual reorientation of the cell-block boundaries toward to RD due to the (cold rolling deformation (Mechanism I, which is the dominant importance) and Hie realignment of boundaries to RD as a result of the shearing introduced by S-bands structure (Mechanism II). During this process a significant number of high angle boundaries (HABs) is created, about, 47% HABs originate from deformation-induced boundaries at 90% reduction. The number of HABs increases and the spacing decreases with the increase of strains. The texture evolves into typical cold rolling deformation texture components of Brass+S+Copper, and the intensity of the texture increases with the increase of strain.
-
A detailed investigation on the grain structure evolution of AA7005 aluminum alloy during hot deformation
Congchang Xu,Hong He,Zhigang Xue,Luoxing Li
DOI: https://doi.org/10.1016/j.matchar.2020.110801
IF: 4.537
2021-01-01
Materials Characterization
Abstract:<p>Obtaining a fine grain structure in Al-Zn-Mg-Cu aluminum alloys with a relatively low alloying content, such as AA7003 and AA7005, is required when they are used for vehicle bodies. However, comprehensive studies on these materials are still needed. In this study, the grain refinement capability of the commercially used AA7005 aluminum alloy through a hot deformation process with relatively a low strain level was investigated. Hot compression tests were performed over a wide range of temperatures, strain rates, and strains (300–500 °C, 0.005–50 s<sup>−1</sup>, and 0.3–1.1, respectively). Diverse types of grain structure evolution, including dynamic recovery (DRV) and dynamic recrystallization (DRX), occurred during the thermomechanical process, and the related mechanisms were analyzed. The results show that DRV occurred regardless of the deformation parameters, while the occurrence of DRX was highly dependent on the parameters. DRX was more favorable at high temperatures, low strain rates, and large strains (e.g., 500 °C, 0.005 s-1, and 1.1, respectively). It formed new high-angle grain boundaries and resulted in fine grains through low-angle grain boundary rotation, high-angle grain boundary migration, and serration mechanisms, which played the primary role in grain refinement. By carefully controlling the deformation parameters, the original coarse grains were successfully refined down from hundreds of micrometers to 8–25 μm. The results indicate that the AA7005 aluminum alloy has a great potential for grain refinement through hot deformation, but special attention should be paid to the parameters.</p>
materials science, multidisciplinary,metallurgy & metallurgical engineering, characterization & testing
-
Local and global tensile deformation behavior of AA7075 sheet material at 673oK and different strain rates
Zhutian Xu,Linfa Peng,Mukesh K. Jain,David Anderson,John Carsley
DOI: https://doi.org/10.1016/j.ijmecsci.2020.106241
IF: 7.3
2021-04-01
International Journal of Mechanical Sciences
Abstract:<p>The constitutive behavior of AA7075 aluminum sheet like other metallic sheet materials is highly temperature and strain rate dependent at elevated temperatures. The uniaxial tensile test is commonly employed to characterize the temperature and strain rate dependent deformation behavior. However, the traditional uniaxial tensile testing and data analysis procedure treats the gauge length of sample as a uniformly-deforming (or homogeneous) region. This may not be accurate due to early localized deformation that is initiated in the gauge region at elevated temperatures. Treating the post-necking hardening behavior of uniaxial tensile samples has been an important task to expand the effective range of true stress-true strain curves at elevated temperatures for metal and alloys. To explore this issue, uniaxial tensile tests on AA7075-F sheets (with as-fabricated temper) were conducted at 673<sup>o</sup> K and four different test speeds. The test procedure also utilized an on-line 2-D digital image correlation (DIC) system method to obtain full-field strain map from the deforming gauge region of the test specimen. Local stress-strain curves at different points along the gauge length were then calculated based on the recorded local in-plane principal stress and strain values from macroscopic force data and current cross-section of the specimen at the chosen points. In this manner, the DIC strain analysis not only enabled the determination of stress-strain response but also the determination of strain rates at specific chosen points along the gauge length. The local stress-strain curves revealed significant deviation at an earlier stage of deformation from the global (or macroscopic) stress-strain response based on the entire gauge region. To estimate the constitutive behavior at different strain rates, a correction method to traditional stress-strain curves was applied to obtain a family of stress-strain curves at different constant strain rates based on the local stress-strain curves. Additionally, the material deformation response in the stress, strain and strain rate space in the 3D Cartesian coordinate system was also constructed for tests conducted at different stretching speeds. Further, a verification test was conducted to see if the above methodology could provide an accurate estimation of local deformation behavior of AA7075 sheet. Furthermore, the above experimental method was also coupled with the Gurson-Tvergarrd-Needleman (GTN) ductile void damage model and incorporated into ABAQUS-Explicit general purpose finite element (FE) analysis code via a VUMAT material subroutine. Good agreement was obtained for both the local and global stress-strain curves between the FE test simulation and experimental stress-strain results. The proposed experimental methodology and the resulting stress-strain-strain rate behavior is believed to be applicable in describing the strain rate-dependent constitutive behavior of AA7075-F sheet metals at 673<sup>o</sup> K.</p>
engineering, mechanical,mechanics
-
Microstructure evolution, constitutive modeling and forming simulation of AA6063 aluminum alloy in hot deformation
Liang Xu,Dayong Zhou,Congchang Xu,Haiyang Zhang,Wenkun Qu,Pengfei Xie,Luoxing Li
DOI: https://doi.org/10.1016/j.mtcomm.2022.105138
IF: 3.8
2022-12-22
Materials Today Communications
Abstract:Hot compression tests of the AA6063 aluminum alloy are conducted on a Gleeble-3500 thermo-mechanical system with different temperatures, strain rates and compression reduction ratios. The microstructure evolution mechanism is analyzed by observing the deformed samples using optical observation techniques. The results show that the microstructure evolution consists of dynamic recovery (DRV) and dynamic recrystallization (DRX). The primary microstructure evolution mechanism shifts from DRV to DRX at the temperature of 648 K. The recrystallized grains emerge firstly along the original grain boundaries, then within the grains, indicating that discontinuous dynamic recrystallization is prone to occur than continuous dynamic recrystallization. A lower strain rate (e.g., 0.01 s −1 ) promotes the occurrences of DRX and grain growth. Geometric dynamic recrystallization occurs only once the strain exceeds a critical value at certain temperature and strain rate. Based on the microstructure evolution mechanism, a set of unified viscoplastic constitutive model considering dislocation density, DRX fraction, and grain size evolution is established to describe the microstructure evolution during and after hot deformation. The established constitutive model is implemented into the finite element (FE) solver DEFORM-3D and uniaxial hot compression simulations are carried out to verify the prediction accuracy of the established model. Numerical procedure is developed to simulate microstructure evolution in a round bar hot extrusion process.
materials science, multidisciplinary
-
Structure and mechanical behavior of ultrafine-grained aluminum-iron alloy stabilized by nanoscaled intermetallic particles
Amandine Duchaussoy,Xavier Sauvage,Kaveh Edalati,Zenji Horita,Gilles Renou,Alexis Deschamps,Frédéric De Geuser
DOI: https://doi.org/10.1016/j.actamat.2019.01.027
IF: 9.4
2019-04-01
Acta Materialia
Abstract:Ultrafine-grained aluminum alloys offer interesting multifunctional properties with a combination of high strength, low electrical resistivity, and low density. However, due to thermally induced grain coarsening, they typically suffer from an intrinsic poor thermal stability. To overcome this drawback, an Al-2%Fe alloy has been selected because of the low solubility of Fe in Al and their highly positive enthalpy of mixing leading to the formation of stable intermetallic particles. The two-phase alloy has been processed by severe plastic deformation to achieve simultaneously submicrometer Al grains and a uniform distribution of nanoscaled intermetallic particles. The influence of the level of deformation on the microstructure has been investigated thanks to transmission electron microscopy and atom probe tomography and it is shown that for the highest strain a partial dissolution of the metastable Al6Fe particle occurred leading to the formation of a Fe super saturated solid solution. The thermal stability, and especially the precipitation of particles from the ultrafine-grained solid solution and the way they pin grain boundaries has been investigated both from static annealing and in-situ transmission electron microscopy experiments. The correlation between microstructural features and microhardness has been established to identify the various strengthening contributions. Finally, it is shown that ultrafine grained high purity Al with less than 0.01 at. % Fe in solid solution could preserve a grain size only 300 nm after 1 h at 250 °C.
materials science, multidisciplinary,metallurgy & metallurgical engineering
-
Dislocation Boundary Structure from Low to Medium Strain of Cold Rolling AA3104 Aluminum Alloy
Zongyong Yao,Guangjie Huang,Andrew Godfrey,Wei Liu,Qing Liu
DOI: https://doi.org/10.1007/s11661-008-9777-x
2009-01-01
Abstract:The evolution of the dislocation boundary structure during the cold rolling of the AA3104 aluminum alloy has been investigated using electron channeling contrast (ECC) imaging and electron backscattered diffraction (EBSD) techniques. The results show that there is a strong correlation between the dislocation boundary structure and the grain orientation. No strong effect of strain level or second-phase particles on the structure-orientation correlation is found. Based on these observations, the microstructures can be classified into one of three types: type A grains, containing two sets of geometrically necessary boundaries (GNBs), type B grains, containing one set of GNBs, and type C grains, consisting of a structure of large dislocation cells. Grains with a type A microstructure have orientations near the copper, brass, and Goss orientations; grains with a type B microstructure are primarily near the S orientation; and grains with a type C microstructure have orientations near the cube orientation. The alignment of the extended dislocation boundaries depends strongly on the grain orientation. In most grains, the boundaries are parallel to the traces of the most active {111} slip planes, as identified by a Schmid factor analysis.