Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Session 4

Chairman: Professor W.B. Hutchinson

Multiscale modelling of the induced plastic anisotropy of IF steel during sheet metal forming
P. Van Houtte and B. Peeters
Department of Metallurgy and Materials Engineering, Katholieke Universiteit Leuven, Belgium

Abstract
Changes of strain path are typically associated with softening/hardening due to the induced plastic anisotropy during the imposed deformation. These effects can significantly influence the strain distribution and may lead to flow localisation and even failure of the material. The present work aims to develop a substantially improved constitutive model for the behaviour for b.c.c. polycrystals during plastic deformation, taking both texture and strain-path induced anisotropy into account. The model should be able to capture the subtle features observed during changing strain paths by appropriately describing the three sources of anisotropy at their length-scale, i.e. the slip processes at a microscopic scale, the dislocation patterns at a mesoscopic scale and the crystallographic deformation texture at a macroscopic scale. A multilevel model has been developed to that purpose. It simulates the development of dislocation patterns in b.c.c. crystals during monotonic or changing strain paths. This dislocation substructure is schematised in terms of cell-block boundaries (CBBs) and cell boundaries (CBs) in addition to the statistically stored dislocations. In order to describe the dislocation patterns in a grain, three local dislocation densities were introduced. The evolution equations for these densities, driven by the current slip activities, were formulated based on Kocks approach. The dislocation ensembles were coupled to the critical resolved shear stresses such that the CBBs reflected latent hardening, the polarity layers associated with these dislocation sheets led to an asymmetry of slip resistance and the CBs together with the statistically stored dislocations were responsible for isotropic hardening. A micro-macro model, in this paper a full-constraints Taylor model, was used to scale-up from the responses of the constituent single crystals to the polycrystalline behaviour. The resulting multiscale work-hardening/softening model for ferritic steels captured the different sources of plastic anisotropy, each at its characteristic length-scale. Validation of this multiscale model is discussed in regard to the prediction of (i) the evolution of dislocation sheets during two-stage strain paths (at the mesoscopic scale) and (ii) the stress-strain curves obtained during these strain-path changes (at the macroscopic scale). At the end of the paper, an example of an application is given, namely a study of the effect of a strain path change on the r-values of sheet material. Keywords: Mechanical properties, anisotropy, dislocation structure, texture, modelling.

Introduction
Changes of strain path occur during many industrial forming processes. Deep-drawing applications are the most obvious example, but even in rolling, the material close to the surface undergoes shears of changing sign. Softening and hardening effects often accompany these strain-path changes [1]. They can lead to flow localisation in the material and even to total failure [2]. Hence there is a need for constitutive models capable of predicting these effects. The underlying causes of the softening/hardening effects at strain path changes are twofold: (i) the crystallographic anisotropy and (ii) the anisotropy caused by dislocation patterns formed during the prestraining [1].

Teodosiu and Hu [3] have proposed a constitutive model which can reproduce the effect of strain path changes on flow stress quite well. However, this model, although inspired by observations of dislocation mechanisms, was formulated at a macroscopic level. It has been possible to combine it with a models for texture induced anisotropy. The two models were coupled at the macroscopic level [2,4]. It would have been more logical to formulate a model for the effect of the dislocation substructure at the level of each individual grain, and then apply a homogenisation procedure in order to obtain the mechanical properties at macroscopic level. Such model would operate on a virtual polycrystal created in computer memory and representing the textured

70

Thermomechanical Processing: Mechanics, Microstructure & Control


polycrystal. In the present paper, we will describe such approach and some of its results. Parts of the theory and the results can also be found in other papers [58]. then simple shear in the opposite sense. The developed microstructure (Figure 2) plays a key role in the behaviour of the material during changing strain paths [1]. Two important relations between the features in the stress-strain curves and the microstructural development in the material exist: (i) Cell block boundaries (CBBs) are formed in the material parallel to the most active slip systems and they give rise to latent hardening. Thus, when new slip systems are activated (by rotation of the crystal or a change in strain path), the dislocation sheets or CBBs act as effective obstacles for the new mobile dislocations (Figures 2e and 2f). Consequently, these dislocation sheets are responsible for the jump in stress response during a cross test. This effect is called the cross effect (c in Figure 1). (ii) A polarity is assigned to each CBB, i.e. an excess of mobile dislocations of one sign are stopped at one side of the CBB, and of the opposite sign on the other side of the CBB (Figure 2b). This polarity will cause an asymmetry of slip resistance and thus, a drop in stress response can be observed during a reverse test (Figures 2c and 2d). This effect is called a Bauschinger effect (b in Figure 1). Based on these microstructural observations Peeters et al. [5] modelled the development of the dislocation substructure at an intragranular scale. The dislocation patterns were idealised by using the concepts of cell block boundaries capturing the latent hardening and polarity and a randomly distributed cell structure. This mesoscopic model was implemented in a full-constraints Taylor model to scale up the response of a polycrystal from

Experimental observations
Figure 1 shows the experimental stress-strain curves of a deep-drawing IF steel for (1) a simple shear test with the shear direction (SD) parallel to the rolling direction (RD) and the shear plane normal (SPN) parallel to the transverse direction (TD), (2) a cross test, 10% true tensile strain in RD, followed by simple shear with SD also parallel to RD and (3) a reverse test, 30% amount of simple shear in RD,

Figure 1. Experimental stress-strain behaviour of an IF steel during different strain path changes. See text for more explanation.

Figure 2. Schematisation of the evolution of dislocation structures during changing strain paths in case of IF steel: (a, b) Thick lines: during a shear test, CBBs develop parallel to primary slip planes. Dashed lines: less important activated slip planes; (c, d) Reverse test: part of the dislocations immobilised at the CBB are released and eventually are immobilised again at other CBBs; (e, f) Cross test (e.g.: tensile deformation after shear): new glide planes tend to become active, but their dislocations have to intersect the existing CBBs first.

71

Thermomechanical Processing: Mechanics, Microstructure & Control


(a) (b)

Figure 3. (a) Substructure resulting from 100% shear deformation. Shear direction as well as the orientation of CBBs are indicated (after Rauch and Schmitt [9]); (b) Schematical representation of the microstructure: cell block boundaries (CBBs) parallel to {110}-planes, the most active slip system; cell boundaries (CBs) of a more random character.

the response of the constituent single crystals. By simulating the microstructural changes inside each grain a complete coupling of the anisotropy due to crystallographic slip, the texture and the dislocation structure is achieved. This crystal plasticity based work-hardening/softening model allowed us to reproduce several characteristic features during changing strain paths [5] and to explain some interesting characteristics in yield loci predictions [8].

Mesoscopical model
The dislocation patterns which develop inside grains are idealised in terms of cell block boundaries and cell structures, in addition to the statistically stored dislocations. Consequently, the model makes use of three dislocation densities: i. The density of immobile dislocations () stored in the cell boundaries (CBs). ii. The density of immobile dislocations (wd) stored in the cell block boundaries (CBBs). This variable reflects the intensity of the dense walls and is the basic microstructural parameter responsible for most of the latent hardening. iii. The directionally movable dislocation density (wp) associated with the cell block boundaries. This variable gives rise to an asymmetry of slip resistance, i.e. the slip resistance in one direction on a slip system will be different from the corresponding value in the reverse direction. A comparison between a TEM micrograph after 100% simple shear deformation [9] and the schematic representation of the internal variables of the model is provided in Figure 3.
Figure 4. Cross slip mechanism for constructing the CBBs [11].

Formation of CBBs Rauch [10] showed that dislocation sheets in a Fe-3%Si single crystal are formed parallel to the planes of highest slip activity. This implies that CBBs could have a {110} or a {112} orientation for BCC metals. However, the available literature suggests that CBBs are parallel to {110}-planes. Due to the 72

fact that (i) TEM evidence on BCC metals deformed at low temperature reveal long screw dislocations, implying that screw dislocations dictate the slip character [10] and (ii) there are no large strain heterogeneities in between the CBBs; the authors believe that double cross-slip mechanism, proposed by Wiedersich [11], is appropriate for explaining the fact that mobile dislocations will contribute to CBBs parallel to their slip plane (Figure 4). During the simulation of the deformation, the model constructs one or two families of CBBs with the following criteria. The most pronounced family of CBBs is generated parallel to the plane of the most active slip system. The second family of CBBs is realised in a similar way parallel to the plane of the second active slip system. A Kocks type law is used to describe the storage and recovery behaviour of dislocations in these cellular structures. Due to the fact that cross slip is assumed to be the most dominant mechanism for the screw dislocations to get immobilised by a CBB, the mean free path of a mobile dislocation, responsible for the construction of CBB i, is scaled with the mean spacing dCBBi between the dislocation sheets i (see Figure 4). Applying Holts law [12] separately to each family of CBBs, the intensity of each CBB i can be calculated as:

Thermomechanical Processing: Mechanics, Microstructure & Control


(1) with b the magnitude of the Burgers vector, i the total slip th rate on the plane of the i most active slip system (e.g. 1 is the total slip rate on the plane of the most active slip system). The immobilisation coefficient Iwd includes the probability of a mobile dislocation being trapped into a CBB. The recovery coefficient Rwd scales with the annihilation length. Note that for stable orientations only 1 slip system will be very active which results in a dislocation structure consisting of elongated cells. In the case that more slip systems are activated the dislocation structure will consist of more equiaxed cells. This is in good agreement with experimental observations [13]. randomised character. It is envisioned that the randomly distributed dislocation mosaic will be responsible for the isotropic hardening in the material. The traditional KocksMecking type law has been formulated for the evolution of this dislocation density: (3) The immobilisation coefficient I includes the probability of a mobile dislocation being trapped into the randomly distributed CBs. The recovery coefficient R scales with the annihilation length. Again, in case of reversal of dislocation flux and in other special cases, different equations are used, see [6] for more details.

Polarity of CBBs Dislocation sheets are always polarised, in the sense that on each side of the sheet there exists an excess of dislocations of the same sign, which is opposite for the two sides of the sheet [14]. This polarity is due to the immobilisation of mobile dislocations moving on slip systems non-coplanar to a CBB. This is reasonable because positive mobile dislocations on a certain slip system are moving in one sense, and the corresponding negative dislocations are moving in the opposite sense. For each currently generated CBB i, the net flux of dislocations from slip systems non-coplanar to that boundary i is calculated as:

The symbol n denotes the number of slip systems, s the slip rate on slip system s (can be positive as well as negative) and usb the unit vector assigned to the slip direction of the system s. Note that there is no contribution for slip activity coplanar with a cell block boundary (usb uiw = 0). A Kocks type law has been used to describe the storage and recovery of these directionally movable dislocations: (2) The immobilisation coefficient Iwp includes the probability of a mobile dislocation being trapped into the polarity layer of a CBB. The recovery coefficient Rwp scales with the annihilation length. This equation implies that directionally movable dislocations are created at the CBB by interactions of mobile dislocations with either CBB dislocations or directionally movable dislocations already stored at the boundary. The recovery takes only place between mobile dislocations and polarised dislocations of opposite sign. In case of reversal of dislocation flux and in other special cases, different equations are used, see [6] for more details.

Behaviour of CBBs and CBs during a change in strain path A more detailed description of the physical behaviour and the modelling of CBBs and CBs during a change in strain path can be found elsewhere [5]. The 2 most important features will be repeated here: i. A change in deformation path or a rotation of a crystal can lead to the activation of new slip systems. These mobile dislocations will construct new CBBs and will destruct old CBBs. In TEM micrographs it seems that the old CBBs remain in place and exist simultaneously with the new ones coming up, but are not storing additional dislocations anymore. These old CBBs disintegrate slowly with further straining by interaction with noncoplanar dislocations. ii. In a reverse test most of the slip systems that were active during the pre-strain remain active, but will operate in the opposite sense. Therefore, the CBBs will not dissolve. However, the dislocations that got stuck at the borders of the CBBs can move very easily away from the CBB. These directionally movable dislocations will be annihilated by dislocations of opposite sign, e.g. CB dislocations. This will result in an increase of the annihilation rate for the randomly distributed CBs. Critical resolved shear stress The critical resolved shear stress on slip system s is taken to be composed of several contributions (1) represents all aspects of the microstructure that are not included in the internal variables selected here (e.g. solid solution effects); (2) the dislocation density in the CBs; (3) the dislocation density in the CBBs, and (4) the directionally movable dislocation density associated with the CBBs. The resultant critical resolved shear stress is obtained from the rule of mixtures (Figure 5, [15]) for the stress in the CBBs and in the cell interiors CBs as: (4)
with

Construction of CBs The CBs, as observed in Figure 3, are clearly of a different nature than the CBBs. Whereas CBBs are rather straight and aligned with active slip systems, the CBs have a more 73

Thermomechanical Processing: Mechanics, Microstructure & Control


Equations 13 and to estimate the internal variables , wd, wp for each grain at the end of the strain increment, including the identification of the slip planes on which the CBBs grow. Equation 4 is then used to estimate the critical resolved shear stresses on each slip system at the time that the first increment ends and the second one begins. These critical resolved shear stresses are then used by the fullconstraints Taylor code during the treatment of the second strain increment. It is necessary to use a Taylor code which is capable of changing all sc from one grain to the next and from one increment to the next. The evolution equations are again integrated explicitly in the second strain increment, and so on. Thus a multilevel model was obtained. It can be used to calculate the evolution of the flow stress, but also to simulate the deformation texture.

where is the volume fraction of CBBs (see Figure 5 for the meaning of w and d).

Multilevel model
To scale up from the single crystal behaviour to the polycrystalline material response the model described above was implemented in a full-constraints Taylor model (see for example [1617]). {110}<111> and {112}<111> slip systems were used. The chosen method simulates a succession of strain increments. All grains of a virtual polycrystal are processed for each increment. In this, the critical resolved shear stress (sc) needs to be known for each slip system in each grain. For the first strain increment, the value valid for annealed material is used (the same value for all slip systems and all grains). This then leads to a value of the slip rates at the beginning of the first strain increment. A very simple explicit integration method is then used, in which it is assumed that the slip rates s remain constant during the strain increment. This then allows to integrate

Calibration of the multilevel model


The extended Taylor model was developed for single-phase b.c.c. polycrystals. The studied material was a Ti-killed IF steel used for deep-drawing applications. The initial texture of this metal was measured by means of X-ray diffraction techniques and discretised into 250 orientations using the statistical method [18]. The extended Taylor model was calibrated to the measured stress-strain responses in five different mechanical tests: i. a simple shear with SD parallel to RD and SPN parallel to TD; ii. 2 cross tests: 10% and 20% true tensile strain in RD, followed by simple shear with SD parallel to the tensile axis and SPN parallel to TD; iii. 2 reverse tests: 15% and 30% amount of simple shear with SD parallel to RD and SPN parallel to TD, then simple shear in the opposite sense. The experimental stress-strain curves and the corresponding simulation results are depicted in Figure 6. The local saturated dislocation densities for the cells, cell block boundaries and the polarity layers, predicted by this model, are 6.71014/m2, 1.31015/m2 and 5.81014/m2, respectively. More details can be found elsewhere [6]. Table 1 shows the values found for the material parameters of the model (see [6] for a complete set of equations).

Figure 5. Sketch of the cell block boundaries (after Mughrabi [15]).

(a)

(b)

Figure 6. Experimental measurements (thick lines) and the corresponding simulation results (thin lines) for the stress-strain behaviour of an IF steel during different two-stage strain path experiments: (a) simple shear test without prestrain and after tensile prestrains of 10 and 20%; (b) reverse tests in simple shear with prestrains of 10%, 20% and 30%. All tests were performed parallel to the rolling direction.

74

Thermomechanical Processing: Mechanics, Microstructure & Control

Mesoscopic validation with reverse tests


A qualitative validation of the extended Taylor model can be achieved by comparing the simulated evolution of the substructure during reverse tests with TEM results by Nesterova et al. [19]. The Kikuchi line technique was used to measure the orientation of the grain observed in the transmission electron microscope. To simulate the substructure in a grain after an imposed deformation it is essential to know the strain history and the initial orientation of the investigated grain. Whereas the strain history is known, the initial orientation is impossible to determine experimentally. Therefore, it was assumed that the extended Taylor model predicts accurately the lattice rotations of the crystals during the imposed deformation. This is reasonable due to the fact that the deformation textures are modelled appropriately. Knowing the lattice rotations of the crystals it was easy to determine the initial orientation of a crystal which would lead after the imposed deformation to the orientation of the grain studied in the transmission electron microscope. In what follows, previously published results [7] are summarised to be able to explain the effect of the stable/unstable character of a grain on the Bauschinger effect observed during reverse tests.

The analysis, performed in [7], reflected the stable character of the studied grain. The calculations performed by the extended Taylor model are depicted in Figure 7a and they confirm the appearance of 1 pronounced family of CBBs parallel to the (11 0) slip plane. The evolution of the directionally movable dislocations associated to the CBBs is also plotted. The negative sign reflects the sense of the polarity vector with respect to the sense of the normal vector on the CBB. Whereas during the pre-strain directionally movable dislocations are immobilised, they are remobilize during reversal of the shear sense. Note that the density of directionally movable dislocations is low.

Unstable orientation In a TEM micrograph (Figure 9) observed in the same sample as investigated in the previous section, two pronounced families of CBBs are apparent: one family parallel to the (011) slip plane and the other parallel to the (101 ) slip plane [19]. It is analysed in [7] that the initial orientation of the studied grain is unstable for the imposed reverse test. Figure 7b depicts the results obtained by the extended Taylor model.

Table 1. Material parameters of the studied IF steel. I Iwd Iwp 2.2 102 9.4 101 5.0 102 10.5 0.9 R [m] Rwd [m] Rwp [m] Rncg [m] Rrev [m] R2 [m] 8.5 1010 2.6 108 3.8 109 2.3 109 1.0 108 1.0 108

Stable orientation A longitudinal plane view TEM micrograph (Figure 8) in a grain of a specimen after a reverse test of 30% simple shear with SD parallel to RD and SPN parallel to TD, then simple shear of 30% in the opposite sense, only showed one pronounced family of CBBs parallel to the (11 0) slip plane and close to the macroscopic shear direction [19].

1 2

(a)

(b)

Figure 7. (a) The evolution of the intensity and polarity of the CBBs in a crystal with stable initial orientation (96.0, 124.5, 54.5) predicted by the extended Taylor model during a reverse test of 30% simple shear with SD parallel to RD and SPN parallel to TD, then simple shear of 30% in the opposite sense; (b) The evolution of the intensity and polarity of the CBBs in a crystal with unstable initial orientation (43.3, 127.2, 43.2) during the same reverse test.

75

Thermomechanical Processing: Mechanics, Microstructure & Control

(a)

(b)

Figure 8. (a) Longitudinal plane view TEM micrograph in a grain of a specimen after a reverse test of 30% simple shear with SD parallel to RD and SPN parallel to TD, then simple shear of 30% in the opposite sense (1 = 30%, 2 = 30%) (after [19]). The grain orientation is (ND)[RD]=(0.66, 0.49, 0.56)[0.39, 0.41, 0.82]. The shear axis is denoted in this figure; (b) The intersection lines of the crystallographic slip planes in this grain and the longitudinal plane.

(a)

(b)

Figure 9. (a) Longitudinal plane view TEM micrograph in a grain of a specimen after a reverse test of 30% simple shear with SD parallel to RD and SPN parallel to TD, then simple shear of 30% in the opposite sense (1 = 30%, 2 = 30%) (after [19]). The grain orientation is (ND)[RD]=(0.53, 0.58, 0.61)[0.25, 0.80, 0.55]. The shear axis is denoted in this figure; (b) The intersection lines of the crystallographic slip planes in this grain and the longitudinal plane

Both the family of CBBs parallel to (101 ) and the family parallel to (011) are predicted correctly. The evolution of the directionally movable dislocations during the reverse test is also plotted. Note that these densities are much higher than in the case of the stable orientation. As a result, the Bauschinger effect is much more pronounced.

Results
Only a few examples are given here. Other results can be found elsewhere [58].

Prediction of stress-strain curves Figure 10a shows the experimental material behaviour during simple shear tests in different angles between the shear direction and the tensile axis, after 20% uniaxial tensile prestrain parallel to RD. The discretised initial 76

texture (250 orientations) and the values of the parameters (Table 1) were provided as input to the multiscale model. Note that the curve for =0 in Figure 10a was among the experimental data used during this identification procedure, whereas the curves for the other angles were not. Hence the predictions by the model for these other angles can be regarded as genuine validation tests. The simulation results, depicted in Figure 10b, are in good agreement with the experimental results for =45 and 90; the result for =135 is less satisfactory. Note that this particular validation test is focused on the question: is it possible to separate the modelling of anisotropy due to crystallographic texture from the modelling of substructure effects? Such separation is one of the basic goals of the present multiscale model. Indeed, the validation of the model for the substructure is done for only one direction in the sheet, after which the model tries to predict the strain path effects in other directions in

Thermomechanical Processing: Mechanics, Microstructure & Control

(a)

(b)

Figure 10. Material behaviour of a prestrained IF steel, 20% tensile strain parallel to RD, during reloading in simple shear plotted as a function of the angle between the shearing direction and the tensile axis: (a) experimental; (b) simulation.

Figure 11. Experimental instantaneous r-values of an as-received Ti-killed IF steel (circles), of a cold-rolled and annealed low-carbon steel after an amount of 30% positive shearing with SD parallel to RD and SPN parallel to TD (triangles) [20]. The values are plotted as a function of the angle between the tensile axis and RD. The full lines (marked Taylor factor in the legend) are r-values which were calculated from the textures using a full-constraints Taylor model [21].

Figure 12. Same as Figure 6, but now the theoretical curves (marked multiscale in the legend) were calculated using the multiscale model and taking the evolution of the dislocation substructure into account [21].

the sheet. When performing the same simulations without considering the substructure, no strain-path-change effects are observed. These results clearly denote the importance of the deformation-induced intragranular microstructure on the plastic anisotropy.

Influence of substructure evolution on instantaneous r-values. The anisotropic behaviour of deep-drawing steels is usually characterised by means of the instantaneous r-value, also called Hill parameter. It is defined as the ratio of the width to thickness plastic strain rate during a uniaxial tensile test. It can in principle be predicted by means of texture models such as the classical full-constraints Taylor model. Such models take the anisotropy due to the texture of the material into account, but they neglect the influence of the intragranular substructure, since the critical resolved shear stresses are assumed to be equal on all slip systems. This indicates that this texture-based method might be acceptable in predicting the initial instantaneous r-values of an almost dislocation-free annealed material. This will be illustrated here for deep-drawing Ti-killed IF steel. Figure 11 shows the r-values as a function of the angle 77

to the rolling direction or a the steel sheet in the as-received condition and after a prestrain of in-plane shear with the shear direction parallel to the rolling direction (Genevois [20]). The figure also shows the r-values derived from the texture using a full-constraints Taylor model (Peeters [21]). The agreement is satisfactory for the steel in as-received condition, but not at all for the steel with the shear prestrain. The calculations were repeated, this time using the multiscale model and taking the prestrain into account. It can be seen in Figure 12 that result is now much better for the steel with prestrain. This improvement is clearly due to the taking into account of the dislocation substructure. The result for the steel as-received (without prestrain) was also a little better (especially at =0, 90 and 180). This small improvement is not related to the substructure, but rather to a more refined algorithm to find the instantaneous r-values as compared the Taylor results given in Figure 11 which have been obtained using standard software. More details are given by Peeters [21].

Summary and conclusions


A procedure is proposed for construction of substructural features, like cell boundaries and cell block boundaries The anisotropy caused by these dislocation ensembles has a large effect on the stress-strain behaviour of b.c.c. polycrystals, especially during changing strain paths. These effects are

Thermomechanical Processing: Mechanics, Microstructure & Control


known as Bauschinger effect and cross effect. It was found that the dislocation structures are modelled adequately during reverse tests for grains with stable/unstable orientation. Reasonable results have also been obtained for the evolution of the directionally movable dislocations associated to the CBBs. It has been shown that crystals with an unstable orientation lead to a larger Bauschinger effect observed during reverse tests than crystals with stable orientation. This model for substructure formation is implemented in a constitutive model for individual grains, based for the rest on the generalised Schmid law. Homogenisation is performed by means of the Taylor theory. Hence a model is obtained for the plastic deformation of polycrystalline materials in which texture effects are automatically implemented. The parameters of the substructure model have been identified by means of experimental stress-strain curves with some strain path changes obtained for polycrystalline materials for which the texture was known. In this, the angle between the principal axes of the applied strain and the rolling direction was not systematically varied. Validation was done on the basis of stress-strain data obtained for other angles. It was found that the influence of this angle on anisotropy effects (due to both texture and substructure) was reasonably well captured by the model. An application was also discussed. It was shown that for prestrained material, the evolution of r-values as a function of the angle with the rolling direction could be adequately predicted by the new model, but not by the classical Taylor model. This means that both texture and substructure must be taken into account for prestrained material. As a final comment, it should be stressed that the multilevel model presented here can probably be adapted easily for high temperature deformation of polycrystalline metallic materials as long as no dynamic recrystallisation takes place. It could then largely contribute to the modelling of the plastic deformation at higher temperatures since it has the potential to describe the global stored energy as a function of strain path, and to compute the evolution rate and relative densities of the different local dislocation densities as a function of crystal orientation and strain history.

References
1. C. Teodosiu, Materials science input to engineering models, Proceedings, Modelling of Plastic Deformation and Its Engineering Applications, Thirteenth Ris International Symposium on Materials Science 1992, S.I. Andersen et al. (eds.) (Ris National Laboratory, Roskilde, Denmark, 1992), 12546. S. Hiwatashi, A. Van Bael, P. Van Houtte and C. Teodosiu, International Journal of Plasticity, 14 (1998), 64769. C. Teodosiu, and Z. Hu, Evolution of the intragranular microstructure at moderate and large strains: modelling and computational significance, Proceedings, Simulation of Materials Processing: Theory, Methods and Applications, Numiform 1995, S.F. Shen and P.R. Dawson (eds.) (Balkema, Rotterdam, 1995), 17382. S. Hiwatashi, A. Van Bael, P. Van Houtte and C. Teodosiu, Computational Materials Science, 9 (1997), 27484. B. Peeters, S.R. Kalidindi, P. Van Houtte and E. Aernoudt, Acta mater., 48 (2000), 212333. B. Peeters, M. Seefeldt, C. Teodosiu, S.R. Kalidindi, P. Van Houtte and E. Aernoudt, Acta Mater., 49 (2001), 160719. B. Peeters, B. Bacroix, C. Teodosiu, P. Van Houtte, and E. Aernoudt, Acta Mater., 49 (2001), 162132. B. Peeters, S.R. Kalidindi, C. Teodosiu, P. Van Houtte, E. Aernoudt, Journal of the Mechanics and Physics of Solids, 50(4) (2002), 783807. E.F. Rauch, J.-H. Schmitt, Mater. Sci. Engng, A113 (1989), 44148. E.F. Rauch, Journal of the Mechanical Behavior of Materials, 4 (1992), 8189. H. Wiedersich, Journal of Applied Physics, 33(3) (1962), 85458 D.L. Holt, J. Appl. Phys., 41 (1970), 3197201. J.V. Fernandes, J.-H. Schmitt, Phil. Mag., A48 (1983), 84170. U.F. Kocks, T. Hasegawa, R.O. Scattergood, Scripta Metall., 14 (1980), 44954. H. Mughrabi, Mat. Sci. and Eng., 85 (1987), 1531. E. Aernoudt, P. Van Houtte and T. Leffers, Deformation and Textures of Metals at Large Strains, in Plastic Deformation and Fracture of Materials (Vol. 6 of Materials Science and Technology: A Comprehensive Treatment, R.W. Cahn, P. Haasen and E.J. Kramer (eds.)), H. Mughrabi (ed.) (VCH, Weinheim, Federal Republic of Germany, 1993) 89136. P. Van Houtte, Textures and Microstructures, 89 (1988), 31350. L.S. Toth, P. Van Houtte, Textures and Microstructures, 19 (1992), 22944. E.V. Nesterova, B. Bacroix, C. Teodosiu, Metallurgical and Materials Transactions A, 32 (2001), 252738. P. Genevois, Thesis, Etude Exprimentale et Modlisation de Comportement Plastique Anisotrope des Tles dAcier en Grandes Transformation (Inst. National Polytechnique, Grenoble, 1992). B. Peeters, PhD Thesis, Multiscale Modelling of the Induced Plastic Anisotropy in IF Steel during Sheet Forming (Department MTM, Katholieke Universiteit Leuven, 2002), ISBN 90-5682-333-7.

2. 3.

4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16.

17. 18. 19. 20.

Acknowledgements
BP acknowledges the financial support through a scholarship of the Fonds voor Wetenschappelijk Onderzoek, Vlaanderen (FWO). Financial support was also provided by the Federal Government of Belgium (DWTC) through the contract IAP P5/8. We thank L. Delannay, M. Seefeldt, E. Aernoudt, S. Kalidindi, E. Rauch and C. Teodosiu for many interesting discussions on this subject.

21.

78

You might also like