Fig 1 - uploaded by Dong-Xu Wen
Content may be subject to copyright.
Initial microstructures of 7075 aluminum alloy in (a) rolling-transverse (R-T) plane and (b) rolling-normal (R-N) plane 

Initial microstructures of 7075 aluminum alloy in (a) rolling-transverse (R-T) plane and (b) rolling-normal (R-N) plane 

Source publication
Article
Full-text available
Three-dimensional crystal plasticity finite element (CPFE) method is used to investigate the hot compressive deformation behaviors of 7075 aluminum alloy. Based on the grain morphology and crystallographic texture of 7075 aluminum alloy, the microstructure-based representative volume element (RVE) model was established by the pole figure inversion...

Contexts in source publication

Context 1
... EBSD system coupled with a Zeiss ULTRA plus scanning electron microscope (SEM) was applied in order to obtain the grain morphology and crystallographic orientation of 7075 aluminum alloy. Since the grain size of 7075 aluminum alloy is relative large, the scan step is chosen to be 2 lm. The rolling-transverse (R-T) and rolling-normal (R-N) planes of specimens were scanned, and the experimental results are shown in Fig. 1. The grain size of 7075 aluminum alloy can be determined as 500 lm 9 250 lm 9 50 lm from the inverse pole figure ...
Context 2
... hot deformation, the severely deformed regions have high stored energy. The nucleation of recrystallized grain is thermally activated process, and the driving force of nucleation is provided by the energy stored during hot deformation. However, the stored energy generated during hot deformation is too low to drive the homogeneous nucleation. Therefore, recrystallization nucleation usually occurs in the regions with local heterogeneities (Ref 69). As shown in Fig. 10, strain localization appears at the triple junction of grains 1, 2, and 3. This region with high strain amplitude may be the potential nucleation site for recrystallized ...
Context 3
... on Strain Distribution Figure 9 shows the effects of grain orientation and misori- entation on the strain field of 7075 aluminum alloy at the deformation degree of 60% in the R-T plane. The grain boundaries are highlighted using black lines. It can be seen that the distribution of strain is inhomogeneous due to different grain orientations. The strain concentration appears at some grain boundaries, resulting in a strain discontinuity at these grain boundaries. However, there is strain continuity across some other grain boundaries. As Raabe figured out (Ref 67), the crystal kinematics significantly affects the strain heteroge- neity between the neighbour grains. The grains with large strains are kinematically soft while the kinematically hard grains have small strains. During hot deformation, the strain path follows soft grains and avoids hard grains. However, the effects of the grain boundary misorientation on the strain distribution are not obvious in the R-T plane since the loading axial is perpendicular to the R-T plane. Figure 10(a) shows the strain distribution of 7075 aluminum alloy at the deformation degree of 60% in the R-N plane. The inhomogeneous strain field can also be found. From Fig. 10(a), a narrow area with relatively large strain value is marked with a quadrangle. Figure 10(b-d) shows the evolution of strain field in the quadrangle with the increase of deformation degree. In the R-N plane, the evolution of the strain field at the grain boundary junction can be easily observed in Fig. 10(b-d). At the deformation degree of 47%, a small area (the red area shown in Fig. 10b) with relatively large strain value appears at the grain boundary junction. As the deformation proceeds, this small area becomes larger and grows from the grain boundary to the grain interior. When the deformation degree is increased to 60%, the strain field with relatively large strain value expands from one grain to the adjacent grains. At the beginning of the hot deformation, the deformation incompatibility appears at the junction of the grain boundaries, resulting in the strain concentration. With the increase of the deformation degree, the strain concentration expands to the grain interior. In general, strain localization at triple junctions is more obvious than that at straight grain boundaries. Similar results from CP simula- tions were reported by Raabe et al. (Ref 67). It was found that the triple junctions might be the nucleation sites for strain localization, and the grain boundary misorientation affects the strain heterogeneity. From Fig. 10(d), it can be seen that the strain field cannot go across the grain boundary between grains 1 and 2, while the strain field can expand from grain 1 to grain 3. The occurrence of this phenomenon is possibly induced by the grain boundary misorientation. The grain boundary angle between grains 1 and 2 is relatively high (36.02°), while that between grains 1 and 3 is a low-angle grain boundary (5.83°). It can be concluded that the strain field can go across the low- angle grain boundaries, while high-angle grain boundaries act as barriers in the R-N plane. Similar simulation results were observed by Sachtleber et al. (Ref 68), who found that the maximum strain appears at high-angle grain boundaries while the effects of low-angle grain boundaries on strain distribution are less ...
Context 4
... on Strain Distribution Figure 9 shows the effects of grain orientation and misori- entation on the strain field of 7075 aluminum alloy at the deformation degree of 60% in the R-T plane. The grain boundaries are highlighted using black lines. It can be seen that the distribution of strain is inhomogeneous due to different grain orientations. The strain concentration appears at some grain boundaries, resulting in a strain discontinuity at these grain boundaries. However, there is strain continuity across some other grain boundaries. As Raabe figured out (Ref 67), the crystal kinematics significantly affects the strain heteroge- neity between the neighbour grains. The grains with large strains are kinematically soft while the kinematically hard grains have small strains. During hot deformation, the strain path follows soft grains and avoids hard grains. However, the effects of the grain boundary misorientation on the strain distribution are not obvious in the R-T plane since the loading axial is perpendicular to the R-T plane. Figure 10(a) shows the strain distribution of 7075 aluminum alloy at the deformation degree of 60% in the R-N plane. The inhomogeneous strain field can also be found. From Fig. 10(a), a narrow area with relatively large strain value is marked with a quadrangle. Figure 10(b-d) shows the evolution of strain field in the quadrangle with the increase of deformation degree. In the R-N plane, the evolution of the strain field at the grain boundary junction can be easily observed in Fig. 10(b-d). At the deformation degree of 47%, a small area (the red area shown in Fig. 10b) with relatively large strain value appears at the grain boundary junction. As the deformation proceeds, this small area becomes larger and grows from the grain boundary to the grain interior. When the deformation degree is increased to 60%, the strain field with relatively large strain value expands from one grain to the adjacent grains. At the beginning of the hot deformation, the deformation incompatibility appears at the junction of the grain boundaries, resulting in the strain concentration. With the increase of the deformation degree, the strain concentration expands to the grain interior. In general, strain localization at triple junctions is more obvious than that at straight grain boundaries. Similar results from CP simula- tions were reported by Raabe et al. (Ref 67). It was found that the triple junctions might be the nucleation sites for strain localization, and the grain boundary misorientation affects the strain heterogeneity. From Fig. 10(d), it can be seen that the strain field cannot go across the grain boundary between grains 1 and 2, while the strain field can expand from grain 1 to grain 3. The occurrence of this phenomenon is possibly induced by the grain boundary misorientation. The grain boundary angle between grains 1 and 2 is relatively high (36.02°), while that between grains 1 and 3 is a low-angle grain boundary (5.83°). It can be concluded that the strain field can go across the low- angle grain boundaries, while high-angle grain boundaries act as barriers in the R-N plane. Similar simulation results were observed by Sachtleber et al. (Ref 68), who found that the maximum strain appears at high-angle grain boundaries while the effects of low-angle grain boundaries on strain distribution are less ...
Context 5
... on Strain Distribution Figure 9 shows the effects of grain orientation and misori- entation on the strain field of 7075 aluminum alloy at the deformation degree of 60% in the R-T plane. The grain boundaries are highlighted using black lines. It can be seen that the distribution of strain is inhomogeneous due to different grain orientations. The strain concentration appears at some grain boundaries, resulting in a strain discontinuity at these grain boundaries. However, there is strain continuity across some other grain boundaries. As Raabe figured out (Ref 67), the crystal kinematics significantly affects the strain heteroge- neity between the neighbour grains. The grains with large strains are kinematically soft while the kinematically hard grains have small strains. During hot deformation, the strain path follows soft grains and avoids hard grains. However, the effects of the grain boundary misorientation on the strain distribution are not obvious in the R-T plane since the loading axial is perpendicular to the R-T plane. Figure 10(a) shows the strain distribution of 7075 aluminum alloy at the deformation degree of 60% in the R-N plane. The inhomogeneous strain field can also be found. From Fig. 10(a), a narrow area with relatively large strain value is marked with a quadrangle. Figure 10(b-d) shows the evolution of strain field in the quadrangle with the increase of deformation degree. In the R-N plane, the evolution of the strain field at the grain boundary junction can be easily observed in Fig. 10(b-d). At the deformation degree of 47%, a small area (the red area shown in Fig. 10b) with relatively large strain value appears at the grain boundary junction. As the deformation proceeds, this small area becomes larger and grows from the grain boundary to the grain interior. When the deformation degree is increased to 60%, the strain field with relatively large strain value expands from one grain to the adjacent grains. At the beginning of the hot deformation, the deformation incompatibility appears at the junction of the grain boundaries, resulting in the strain concentration. With the increase of the deformation degree, the strain concentration expands to the grain interior. In general, strain localization at triple junctions is more obvious than that at straight grain boundaries. Similar results from CP simula- tions were reported by Raabe et al. (Ref 67). It was found that the triple junctions might be the nucleation sites for strain localization, and the grain boundary misorientation affects the strain heterogeneity. From Fig. 10(d), it can be seen that the strain field cannot go across the grain boundary between grains 1 and 2, while the strain field can expand from grain 1 to grain 3. The occurrence of this phenomenon is possibly induced by the grain boundary misorientation. The grain boundary angle between grains 1 and 2 is relatively high (36.02°), while that between grains 1 and 3 is a low-angle grain boundary (5.83°). It can be concluded that the strain field can go across the low- angle grain boundaries, while high-angle grain boundaries act as barriers in the R-N plane. Similar simulation results were observed by Sachtleber et al. (Ref 68), who found that the maximum strain appears at high-angle grain boundaries while the effects of low-angle grain boundaries on strain distribution are less ...
Context 6
... on Strain Distribution Figure 9 shows the effects of grain orientation and misori- entation on the strain field of 7075 aluminum alloy at the deformation degree of 60% in the R-T plane. The grain boundaries are highlighted using black lines. It can be seen that the distribution of strain is inhomogeneous due to different grain orientations. The strain concentration appears at some grain boundaries, resulting in a strain discontinuity at these grain boundaries. However, there is strain continuity across some other grain boundaries. As Raabe figured out (Ref 67), the crystal kinematics significantly affects the strain heteroge- neity between the neighbour grains. The grains with large strains are kinematically soft while the kinematically hard grains have small strains. During hot deformation, the strain path follows soft grains and avoids hard grains. However, the effects of the grain boundary misorientation on the strain distribution are not obvious in the R-T plane since the loading axial is perpendicular to the R-T plane. Figure 10(a) shows the strain distribution of 7075 aluminum alloy at the deformation degree of 60% in the R-N plane. The inhomogeneous strain field can also be found. From Fig. 10(a), a narrow area with relatively large strain value is marked with a quadrangle. Figure 10(b-d) shows the evolution of strain field in the quadrangle with the increase of deformation degree. In the R-N plane, the evolution of the strain field at the grain boundary junction can be easily observed in Fig. 10(b-d). At the deformation degree of 47%, a small area (the red area shown in Fig. 10b) with relatively large strain value appears at the grain boundary junction. As the deformation proceeds, this small area becomes larger and grows from the grain boundary to the grain interior. When the deformation degree is increased to 60%, the strain field with relatively large strain value expands from one grain to the adjacent grains. At the beginning of the hot deformation, the deformation incompatibility appears at the junction of the grain boundaries, resulting in the strain concentration. With the increase of the deformation degree, the strain concentration expands to the grain interior. In general, strain localization at triple junctions is more obvious than that at straight grain boundaries. Similar results from CP simula- tions were reported by Raabe et al. (Ref 67). It was found that the triple junctions might be the nucleation sites for strain localization, and the grain boundary misorientation affects the strain heterogeneity. From Fig. 10(d), it can be seen that the strain field cannot go across the grain boundary between grains 1 and 2, while the strain field can expand from grain 1 to grain 3. The occurrence of this phenomenon is possibly induced by the grain boundary misorientation. The grain boundary angle between grains 1 and 2 is relatively high (36.02°), while that between grains 1 and 3 is a low-angle grain boundary (5.83°). It can be concluded that the strain field can go across the low- angle grain boundaries, while high-angle grain boundaries act as barriers in the R-N plane. Similar simulation results were observed by Sachtleber et al. (Ref 68), who found that the maximum strain appears at high-angle grain boundaries while the effects of low-angle grain boundaries on strain distribution are less ...
Context 7
... on Strain Distribution Figure 9 shows the effects of grain orientation and misori- entation on the strain field of 7075 aluminum alloy at the deformation degree of 60% in the R-T plane. The grain boundaries are highlighted using black lines. It can be seen that the distribution of strain is inhomogeneous due to different grain orientations. The strain concentration appears at some grain boundaries, resulting in a strain discontinuity at these grain boundaries. However, there is strain continuity across some other grain boundaries. As Raabe figured out (Ref 67), the crystal kinematics significantly affects the strain heteroge- neity between the neighbour grains. The grains with large strains are kinematically soft while the kinematically hard grains have small strains. During hot deformation, the strain path follows soft grains and avoids hard grains. However, the effects of the grain boundary misorientation on the strain distribution are not obvious in the R-T plane since the loading axial is perpendicular to the R-T plane. Figure 10(a) shows the strain distribution of 7075 aluminum alloy at the deformation degree of 60% in the R-N plane. The inhomogeneous strain field can also be found. From Fig. 10(a), a narrow area with relatively large strain value is marked with a quadrangle. Figure 10(b-d) shows the evolution of strain field in the quadrangle with the increase of deformation degree. In the R-N plane, the evolution of the strain field at the grain boundary junction can be easily observed in Fig. 10(b-d). At the deformation degree of 47%, a small area (the red area shown in Fig. 10b) with relatively large strain value appears at the grain boundary junction. As the deformation proceeds, this small area becomes larger and grows from the grain boundary to the grain interior. When the deformation degree is increased to 60%, the strain field with relatively large strain value expands from one grain to the adjacent grains. At the beginning of the hot deformation, the deformation incompatibility appears at the junction of the grain boundaries, resulting in the strain concentration. With the increase of the deformation degree, the strain concentration expands to the grain interior. In general, strain localization at triple junctions is more obvious than that at straight grain boundaries. Similar results from CP simula- tions were reported by Raabe et al. (Ref 67). It was found that the triple junctions might be the nucleation sites for strain localization, and the grain boundary misorientation affects the strain heterogeneity. From Fig. 10(d), it can be seen that the strain field cannot go across the grain boundary between grains 1 and 2, while the strain field can expand from grain 1 to grain 3. The occurrence of this phenomenon is possibly induced by the grain boundary misorientation. The grain boundary angle between grains 1 and 2 is relatively high (36.02°), while that between grains 1 and 3 is a low-angle grain boundary (5.83°). It can be concluded that the strain field can go across the low- angle grain boundaries, while high-angle grain boundaries act as barriers in the R-N plane. Similar simulation results were observed by Sachtleber et al. (Ref 68), who found that the maximum strain appears at high-angle grain boundaries while the effects of low-angle grain boundaries on strain distribution are less ...
Context 8
... on Strain Distribution Figure 9 shows the effects of grain orientation and misori- entation on the strain field of 7075 aluminum alloy at the deformation degree of 60% in the R-T plane. The grain boundaries are highlighted using black lines. It can be seen that the distribution of strain is inhomogeneous due to different grain orientations. The strain concentration appears at some grain boundaries, resulting in a strain discontinuity at these grain boundaries. However, there is strain continuity across some other grain boundaries. As Raabe figured out (Ref 67), the crystal kinematics significantly affects the strain heteroge- neity between the neighbour grains. The grains with large strains are kinematically soft while the kinematically hard grains have small strains. During hot deformation, the strain path follows soft grains and avoids hard grains. However, the effects of the grain boundary misorientation on the strain distribution are not obvious in the R-T plane since the loading axial is perpendicular to the R-T plane. Figure 10(a) shows the strain distribution of 7075 aluminum alloy at the deformation degree of 60% in the R-N plane. The inhomogeneous strain field can also be found. From Fig. 10(a), a narrow area with relatively large strain value is marked with a quadrangle. Figure 10(b-d) shows the evolution of strain field in the quadrangle with the increase of deformation degree. In the R-N plane, the evolution of the strain field at the grain boundary junction can be easily observed in Fig. 10(b-d). At the deformation degree of 47%, a small area (the red area shown in Fig. 10b) with relatively large strain value appears at the grain boundary junction. As the deformation proceeds, this small area becomes larger and grows from the grain boundary to the grain interior. When the deformation degree is increased to 60%, the strain field with relatively large strain value expands from one grain to the adjacent grains. At the beginning of the hot deformation, the deformation incompatibility appears at the junction of the grain boundaries, resulting in the strain concentration. With the increase of the deformation degree, the strain concentration expands to the grain interior. In general, strain localization at triple junctions is more obvious than that at straight grain boundaries. Similar results from CP simula- tions were reported by Raabe et al. (Ref 67). It was found that the triple junctions might be the nucleation sites for strain localization, and the grain boundary misorientation affects the strain heterogeneity. From Fig. 10(d), it can be seen that the strain field cannot go across the grain boundary between grains 1 and 2, while the strain field can expand from grain 1 to grain 3. The occurrence of this phenomenon is possibly induced by the grain boundary misorientation. The grain boundary angle between grains 1 and 2 is relatively high (36.02°), while that between grains 1 and 3 is a low-angle grain boundary (5.83°). It can be concluded that the strain field can go across the low- angle grain boundaries, while high-angle grain boundaries act as barriers in the R-N plane. Similar simulation results were observed by Sachtleber et al. (Ref 68), who found that the maximum strain appears at high-angle grain boundaries while the effects of low-angle grain boundaries on strain distribution are less ...
Context 9
... the texture tends to be stable, Fig. 11 shows only the 111 h i pole figures of grains 1, 2, and 3 after the compressive compression. It can be found that the pole figures of grains 1 and 3 are almost the same, and different from that of grain 2. It reveals that the grain boundary between grains 1 and 2 is still a high-angle grain boundary, while that between grains 1 and 3 is a low-angle grain boundary. The initial orientation of each element in one grain is the same, and the texture shown in pole figure is concentrated. As can be seen from Fig. 11, the texture is divergent after the deformation, which indicates that the deformation heterogeneity occurs in every grain (shown in Fig. 10). Zhao et al. (Ref 70) also found that the strain heterogeneity is significantly affected by the microtexture through experiments and 3D CP simulations. This is due to the separation of in-grain kinematics and grain interaction, as discussed by Raabe et al. (Ref ...
Context 10
... the texture tends to be stable, Fig. 11 shows only the 111 h i pole figures of grains 1, 2, and 3 after the compressive compression. It can be found that the pole figures of grains 1 and 3 are almost the same, and different from that of grain 2. It reveals that the grain boundary between grains 1 and 2 is still a high-angle grain boundary, while that between grains 1 and 3 is a low-angle grain boundary. The initial orientation of each element in one grain is the same, and the texture shown in pole figure is concentrated. As can be seen from Fig. 11, the texture is divergent after the deformation, which indicates that the deformation heterogeneity occurs in every grain (shown in Fig. 10). Zhao et al. (Ref 70) also found that the strain heterogeneity is significantly affected by the microtexture through experiments and 3D CP simulations. This is due to the separation of in-grain kinematics and grain interaction, as discussed by Raabe et al. (Ref ...
Context 11
... the texture tends to be stable, Fig. 11 shows only the 111 h i pole figures of grains 1, 2, and 3 after the compressive compression. It can be found that the pole figures of grains 1 and 3 are almost the same, and different from that of grain 2. It reveals that the grain boundary between grains 1 and 2 is still a high-angle grain boundary, while that between grains 1 and 3 is a low-angle grain boundary. The initial orientation of each element in one grain is the same, and the texture shown in pole figure is concentrated. As can be seen from Fig. 11, the texture is divergent after the deformation, which indicates that the deformation heterogeneity occurs in every grain (shown in Fig. 10). Zhao et al. (Ref 70) also found that the strain heterogeneity is significantly affected by the microtexture through experiments and 3D CP simulations. This is due to the separation of in-grain kinematics and grain interaction, as discussed by Raabe et al. (Ref ...
Context 12
... M 1 and M 2 are the orientation matrices for the initial and current orientations, respectively. The misorientation DH can be computed from the misorientation matrix A mis , and DH is formulated by Eq 11. There are 24 crystallographically equivalent misorientations considering the symmetry of the FCC crystal. The minimum of all the misorientations is adopted as the final misorientation for each element. Figure 12 shows the misorientation distribution of 7075 aluminum alloy after the hot deformation. The value of misorientation can be used to represent the rotation degree of each crystal. From Fig. 12, it can be found that the distribution of misorientation is inhomogeneous and the maximum misori- entation angle is 52.76°. The maximum misorientation appears at the edges of the RVE model, and most of the misorientations are less than 30°. The relatively homogeneous distribution of misorientation within the grain is observed, and misorientation continuity or discontinuity at the grain boundaries can be found. Also, the effects of grain boundary on the misorientation distribution are obvious in the R-N plane. As shown in Fig. 12(b), the misorientation of the grain boundary between grains 4 and 5 is 6.26°, while that between grains 5 and 6 is 55.68°. It can also be concluded that the misorientation continuity is present across the low-angle grain boundary, while the misorientation discontinuity appears at the high-angle grain boundary. Therefore, it can be concluded that the grain rotation incompatibility can occur at the high-angle grain boundary, and the low-angle grain boundary has little effect on the misorientation distribution during the hot-compressive ...
Context 13
... M 1 and M 2 are the orientation matrices for the initial and current orientations, respectively. The misorientation DH can be computed from the misorientation matrix A mis , and DH is formulated by Eq 11. There are 24 crystallographically equivalent misorientations considering the symmetry of the FCC crystal. The minimum of all the misorientations is adopted as the final misorientation for each element. Figure 12 shows the misorientation distribution of 7075 aluminum alloy after the hot deformation. The value of misorientation can be used to represent the rotation degree of each crystal. From Fig. 12, it can be found that the distribution of misorientation is inhomogeneous and the maximum misori- entation angle is 52.76°. The maximum misorientation appears at the edges of the RVE model, and most of the misorientations are less than 30°. The relatively homogeneous distribution of misorientation within the grain is observed, and misorientation continuity or discontinuity at the grain boundaries can be found. Also, the effects of grain boundary on the misorientation distribution are obvious in the R-N plane. As shown in Fig. 12(b), the misorientation of the grain boundary between grains 4 and 5 is 6.26°, while that between grains 5 and 6 is 55.68°. It can also be concluded that the misorientation continuity is present across the low-angle grain boundary, while the misorientation discontinuity appears at the high-angle grain boundary. Therefore, it can be concluded that the grain rotation incompatibility can occur at the high-angle grain boundary, and the low-angle grain boundary has little effect on the misorientation distribution during the hot-compressive ...
Context 14
... M 1 and M 2 are the orientation matrices for the initial and current orientations, respectively. The misorientation DH can be computed from the misorientation matrix A mis , and DH is formulated by Eq 11. There are 24 crystallographically equivalent misorientations considering the symmetry of the FCC crystal. The minimum of all the misorientations is adopted as the final misorientation for each element. Figure 12 shows the misorientation distribution of 7075 aluminum alloy after the hot deformation. The value of misorientation can be used to represent the rotation degree of each crystal. From Fig. 12, it can be found that the distribution of misorientation is inhomogeneous and the maximum misori- entation angle is 52.76°. The maximum misorientation appears at the edges of the RVE model, and most of the misorientations are less than 30°. The relatively homogeneous distribution of misorientation within the grain is observed, and misorientation continuity or discontinuity at the grain boundaries can be found. Also, the effects of grain boundary on the misorientation distribution are obvious in the R-N plane. As shown in Fig. 12(b), the misorientation of the grain boundary between grains 4 and 5 is 6.26°, while that between grains 5 and 6 is 55.68°. It can also be concluded that the misorientation continuity is present across the low-angle grain boundary, while the misorientation discontinuity appears at the high-angle grain boundary. Therefore, it can be concluded that the grain rotation incompatibility can occur at the high-angle grain boundary, and the low-angle grain boundary has little effect on the misorientation distribution during the hot-compressive ...

Similar publications

Article
Full-text available
The power dissipation maps of Ti-25Al-15Nb alloy were constructed by using the compression test data. A method is proposed to predict the distribution and variation of power dissipation coefficient in hot forging process using both the dynamic material model and finite element simulation. Using the proposed method, the change characteristics of the...

Citations

... At the grain level, in view of the spatial local strain field distribution inside polycrystal and deformation texture, 3D crystal plasticity finite element method (CPFEM) framework was introduced by Li. et al (Li et al., 2015) and Tu (Tu et al., 2019). It was concluded that strain heterogeneity is more severe along the high angle grain boundaries rather than along low angle grain boundaries. ...
Article
Elevated temperature forming of AA7075 aluminum alloy is of considerable interest in the automotive industry due to its potential for light weighting and several other environmental benefits. This study presents elevated temperature uniaxial tensile deformation and damage behavior of AA7075-W sheet through a set of experiments at the microstructural scale. A multi-scale experimental methodology consisting of in-situ SEM tensile testing, use of micro digital image correlation (µ-DIC) method for obtaining strain field at the microstructural scale and post-test X-ray computed tomography was adopted with temperature held constant at 300°C and 400°C. The deformation temperature affected the precipitate formation and their distribution and subsequently the precipitate-induced void damage development in the material. At 300°C, precipitates as well as precipitate free zones were formed at the grain boundaries during heating and following deformation. Intergranular void initiation and catastrophic void coalescence with relatively small void growth occurred along the grain boundaries during plastic straining resulting in intergranular fracture and poor ductility at this temperature. In contrast, at 400°C, fewer intergranular precipitates were formed during heating and deformation resulting in delayed void nucleation during plastic straining. The diffuse necking occurred earlier with significant void growth occurring during the long regime of diffuse necking with very little evidence of void coalescence leading to considerably higher ductility at fracture. The final fracture occurred due to transgranular microcrack formation and propagation at this temperature.
... A given number of grain orientations can be uniquely generated with this method, which has been implemented as part of the open-source crystallographic toolbox MTEX [27]. Li et al. [28] used MTEX to discretize ODF to obtain a given amount of grain orientations, which were then assigned to the RVE model of 7075 aluminum alloy. The process of building the complete model and simulating is shown in Figure 4. ...
Article
Full-text available
The deformation process of metal foils is usually under a complex stress status, and the size effect has an obvious influence on the microforming process. To study the effect of grain orientation and grain size distribution on the yield loci evolution of SUS304 stainless steel foils, three representative volume element (RVE) models were built based on the open source tools NEPER and MTEX In addition, the yield loci with different grain sizes are obtained by simulation with Duisseldorf Advanced Material Simulation Kit (DAMASK) under different proportional loading conditions. The initial yield loci show a remarkable difference in shape and size, mainly caused by the distinct texture characteristics. By comparing the crystal plasticity simulation with the experimental results, the model with normal grain size distribution and initial texture based on Electron Back-scattered Diffraction (EBSD) data can more accurately describe the influence of the size effect on the shape and size of yield loci, which is the result of the interaction of grain size distribution and texture. However, the enhancement of grain deformation coordination will weaken the impact of the size effect on yield loci shape if the grain size distribution is more uniform.
... Similarly, by generating and operating on these newly created ODFs, the corresponding fourpeak texture and annular texture can be depicted. Lastly, these representative ODFs can be discretized by using the code ''calcOrientations'' in MATLAB Table 1 Material parameters of all deformation modes in VPSC modeling (MPa) [24] Mode software to obtain the texture input containing 5000 grain orientations, as shown in Fig. 1 [26,27]. The axis of X, Y and Z here indicate RD, TD and ND, respectively. ...
Article
Full-text available
Effects of texture on micromechanics, texture evolution and macromechanics of AZ31 magnesium alloy sheets with tailored textures during plastic deformation are thoroughly investigated by visco-plastic self-consistent (VPSC) analysis. These simulated results confirm that as for in-plane tension (IPT), activities of basal < a > slip and \(\{ 10\overline{1}2\}\) extension twin (ET) are mainly responsible for the observed texture characteristics of all deformed samples. In addition, activities of prismatic < a > slip and \(\{ 10\overline{1}2\}\) ET in samples with bimodal texture are highly related to the loading direction. With regard to in-plane shear (IPS), sample with weak basal texture possesses the largest activity of prismatic < a > slip. Moreover, basal < a > slip and prismatic < a > slip collectively contribute to the rotation of tilted basal poles in (0002) pole figure, whereas pyramidal < c + a > slip and \(\{ 10\overline{1}2\}\) ET are advantageous to the rotation of tilted basal poles away from normal direction (ND) with increasing plastic strain. In terms of plane strain compression (PSC), \(\{ 10\overline{1}2\}\) ET and prismatic < a > slip largely participate in plastic deformation of samples with four-peak texture and annular (ring-shaped) texture, namely, \(\{ 10\overline{1}2\}\) ET leads to the rotation of tilted basal poles toward ND during PSC and prismatic < a > slip causes the formation of a small annular texture component after a relatively large plastic strain.
... Recently, crystal plasticity (CP) models have been extensively applied to predict texture evolution during plastic deformation process of metal materials. Li et al. established a microstructure-based representative volume element model and used crystal plasticity finite element method (CPFEM) to investigate the microstructural evolution of 7075 aluminium alloy and found that the predicted deformation texture agreed well with the experimental observation [2]. Jung et al. investigated texture evolution during the cold rolling process of aluminium alloy AA1050-O using CPFEM [3]. ...
Article
Accurate prediction of the texture evolution of aluminium alloys during hot extrusion process plays an important role in the die design and determination of process parameters. In this work, the deformation texture of aluminium alloys under three-dimensional deformation path are investigated numerically and experimentally. Firstly, the extrusion process of a helical profile is simulated with HyperXtrude and the deformation history under three-dimensional deformation path of the profile is extracted. Then the deformation modes in the different positions on the profile are analyzed and the velocity gradients are imported respectively into FC-Taylor, ALAMEL and ALAMEL III models to predict the extrusion deformation textures. Finally, a specially-designed helical extrusion die is manufactured and the extrusion experiment are carried out. The textures are analyzed by EBSD to verify the predicted results by three crystal plasticity models. The results show that there are significant differences in the deformation modes from the centre to surface of the extruded profile and the three crystal plasticity models differ greatly in the prediction accuracy of textures. For the deformation texture of the helical profile, the FC-Taylor model gives better prediction at the centre position and the ALAMEL-type model predicts better at the surface position of the profile.
... C 11 , C 12 , and C 44 are the elastic constants. _ c a and n are the reference strain rate and the rate sensitivity exponent for the slip system a; h 0 , s s, and s 0 are the initial hardening modulus, stage I stress, and initial yield stress, respectively; q is the ratio of latent over self-hardening moduli within the same set of slip systems ( Ref 29,30). ...
Article
To understand the influence of recrystallization on the low-cycle fatigue behavior of directionally solidified nickel-based DZ4 superalloy, crystal plasticity-based finite element (CPFE) analysis was carried out to study the cyclic plastic behaviors. Using the design of experiments (DOE) technique, a series of CPFE simulations were designed and performed for DZ4 superalloy with recrystallized grains to examine the cyclic plastic shear deformation at the carbide/matrix interface, at the recrystallized grain boundary/matrix, and in the matrix. Seven influence factors, i.e., applied strain, strain ratio, carbide modulus, dwell time, misorientation angle, carbide aspect ratio, and carbide size, were considered, and the effects of which were quantitatively evaluated by the maximum accumulated plastic shear strains via CPFE-based DOE analysis. The results showed that carbide and recrystallized grain boundary had competitively significant effects on fatigue crack initiation. Applied strain was the most influential parameter among all the studied parameters.
... The texture evolution and mechanical response of polycrystalline materials at the micro-scale can be analyzed with crystal plasticity (CP) [6,7,8,9,10], which leads to the modeling for the micro-mechanisms of plastic deformation using elasto-viscoplastic constitutive laws [11] for various slip systems. With the CP computational models, the texture evolution and the mechanical response of the material during deformation can be simulated using parameters such as slip system strengths and hardening rates. ...
Article
Full-text available
The present study addresses the multi-scale computational modeling of a lightweight Aluminum-Lithium (Al-Li) 2070 alloy. The Al-Li alloys display significant anisotropy in material properties because of their strong crystallographic texture. To understand the relationships between processing, microstructural textures at different material points and tailored material properties, a multi-scale simulation is performed by controlling the texture evolution during deformation. To achieve the multi-scale framework, a crystal plasticity model based on a one-point probability descriptor, Orientation Distribution Function (ODF), is implemented to study the texture evolution. Next, a two-way coupled multi-scale model is developed, where the deformation gradient at the macro-scale integration points is passed to the micro- scale ODF model and the homogenized stress tensor at the micro-scale is passed back to the macro-scale model. A gradient-based optimization scheme which incorporates the multi-scale continuum sensitivity method is utilized to calibrate the slip system parameters of the alloy using the available experimental data. Next, the multi-scale simulations are performed for compression and tension using the calibrated crystal plasticity model, and the texture data is compared to the experiments. With the presented multi-scale modeling scheme, we achieve the location-specific texture predictions for a new generation Al-Li alloy for different deformation processes.
... Moreover, the initial orientations of individual grains are obtained by discretizing the orientation distribution function (ODF) of EBSD data. Further details of orientation discretization and assignment can be found in [27]. The set of constitutive parameters used in the present study is shown in Table 1, where the elastic constants of NiTi SMA are obtained from the literature [9]. ...
Article
Full-text available
Numerical modeling of microstructure evolution in various regions during uniaxial compression and canning compression of NiTi shape memory alloy (SMA) are studied through combined macroscopic and microscopic finite element simulation in order to investigate plastic deformation of NiTi SMA at 400 °C. In this approach, the macroscale material behavior is modeled with a relatively coarse finite element mesh, and then the corresponding deformation history in some selected regions in this mesh is extracted by the sub-model technique of finite element code ABAQUS and subsequently used as boundary conditions for the microscale simulation by means of crystal plasticity finite element method (CPFEM). Simulation results show that NiTi SMA exhibits an inhomogeneous plastic deformation at the microscale. Moreover, regions that suffered canning compression sustain more homogeneous plastic deformation by comparison with the corresponding regions subjected to uniaxial compression. The mitigation of inhomogeneous plastic deformation contributes to reducing the statistically stored dislocation (SSD) density in polycrystalline aggregation and also to reducing the difference of stress level in various regions of deformed NiTi SMA sample, and therefore sustaining large plastic deformation in the canning compression process.
... Accounting for the physical background of material, the theory of dislocation density, thermodynamics, and crystal plasticity are successively introduced, and some physically based constitutive models were reported. [18][19][20] The detailed background of dislocation density-based model is introduced in Sec. II. ...
Article
Generally, the obvious work hardening, dynamic recrystallization (DRX), and dynamic recovery behaviors can be found during hot deformation of Ni-based superalloys. In the present study, the classical dislocation density theory is improved by introducing a new dislocation annihilation item to represent the influences of DRX on dislocation density evolution for a Ni-based superalloy. Based on the improved dislocation density theory, the peak strain corresponding to peak stress and the critical strain for initiating DRX can be determined, and the improved DRX kinetics equations and grain size evolution models are developed. The physical framework and algorithmic idea of the improved dislocation density theory are clarified. Moreover, the deformed microstructures are characterized and quantitatively correlated to validate the improved dislocation density theory. It is found that the improved dislocation density-based models can precisely characterize hot deformation and DRX behaviors for the studied superalloy under the tested conditions.