• Aucun résultat trouvé

Microstructures, textures and mechanical properties after large strain

N/A
N/A
Protected

Academic year: 2021

Partager "Microstructures, textures and mechanical properties after large strain"

Copied!
14
0
0

Texte intégral

(1)

HAL Id: jpa-00245799

https://hal.archives-ouvertes.fr/jpa-00245799

Submitted on 1 Jan 1988

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

N. Hansen, T. Leffers

To cite this version:

N. Hansen, T. Leffers. Microstructures, textures and mechanical properties after large strain.

Revue de Physique Appliquée, Société française de physique / EDP, 1988, 23 (4), pp.519-531.

�10.1051/rphysap:01988002304051900�. �jpa-00245799�

(2)

Microstructures, textures and mechanical properties after large strain

N. Hansen and T. Leffers

Metallurgy Department, Risø National Laboratory, 4000 Roskilde, Denmark

(Reçu le 26 mai 1987, accepté le 28 septembre 1987)

Résumé. - On passe en revue les propriétés mécaniques, la texture et la microstructure des matériaux

polycristallins. Les modèles existants pour la déformation sont discutés et les propriétés mécani-

ques sont corrélées avec la microstructure. Enfin, l’anisotropie et les relations constitutives sont discutées.

Abstract. - Microstructure, texture and mechanical properties of polycrystalline materials after medium and large strains are reviewed. The existing deformation models are discussed on the basis of the evidence presented, and mechanical properties and microstructure are correlated. Finally, anisotropy and constitutive relations are discussed briefly.

61.50J - 61.70L - 62.20 - 62.20F

CONTENTS 1. Introduction

2. Deformation microstructures 2.1. Dislocation cells 2.2. Bands

2.3. Surface observations

2.4. Effect of metallurgical parameters 3. Deformation textures

3.1. Rolling textures

3.2. Effect of metallurgical parameters 4. Deformation models

4.1. Models and microscopical observations 4.2. Models and texture

4.3. Improved models 5. Mechanical properties

5.1. Flow stress-strain behaviour 5.2. Flow stress-texture (M factors) 5.3. Flow stress-microstructural evolution 5.4. Flow stress-cell size

5.5. Effect of metallurgical parameters 5.6. Anisotropy

6. Constitutive relations

Concluding remarks

1. INTRODUCTION

Deformation of metals and alloys is accompanied by changes in microstructure and texture. These

changes can provide information about the mechan- isms of plastie flow and about strain-induced

changes in the mechanical properties. The present paper will deal with medium and large deformations of polycrystalline materials at low temperatures (where climb can be neglected). A characteristie behaviour is examplified by f.c.e. metals (alu-

minium and copper), and it is discussed how this behaviour is modified by metallurgical parameters such as grain size and volume fraction and size of

hard particles. The microstructural and textural information is related to the flow stress-strain behaviour, and the relationship between the flow stress and various microstructural parameters is described. Plastic anisotropy is described briefly,

and finally the possibility of developing consti-

tutive relations on theoretical basis is diseussed.

2. DEFORMATION MICROSTRUCTURES

The dislocation microstructures in deformed ma- terials have been studied extensively in order to understand the deformation mechanisms and to derive relations between the flow stress and micro- structural parameters. In general the dislocation structures develop with increasing strain from

tangled dislocations to well-defined dislocation arrays forming a regular cell structure. Superim- posed on the cell structure, a banded dislocation structure is frequently observed. The various types of dislocation arrangements will be summarized in the following. In section 4 these mieroscopical observations (combined with texture observations) will be discussed in relation to the existing de-

formation models.

2.1. Dislocation cells.

The formation of dislocation cells has been dis- eussed on the basis of different mechanisms of which the energy minimization principle recently

has been treated in detail [1]. It follows from this principle that a uniform or statistical dis- tribution of dislocations is unstable since the dislocation energy can be reduced when the dislo- cations cluster into stress-screening arrays. The degree of this réduction depends on the availabili- ty of slip systems because an inerease in the number of Burgers vectors permits the formation of arrays of lower energy. Mobility of dislocations is another important factor, whieh may be influenced

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/rphysap:01988002304051900

(3)

by cross slip. Clustering of dislocations leads to

a structure consisting of volume elements almost free of dislocations separated by densely packed dislocation arrays, and this is what one observes at medium and high strains. As an example figure 1 shows the cell structure of aluminium cold drawn 80

%.

Fig. 1. TEM micrograph of 80% cold-drawn 99.998% Al, longitudinal section. From [2].

An alternative mechanism for the formation of cell structures has been suggested [3] on the basis of the existence of ’hard areas’ in the crystals which

are impenetrable for dislocations. Three-dimension- al boundaries may then form and there will be a strong tendency that dislocation are placed on each side of the boundary according to their sign.

Thereby a bipolar wall is formed, which contains a

high local internal stress. This is in contrast to the virtually stress free walls formed according to the energy minimization principle.

For a regular cell structure as shown in Fig. 1 characteristic features are the cell size and the misorientation across the cell walls. The dislo- cation arrangement in the cell walls may be quite simple as in tilt and twist boundaries. However, in most cases especially at medium and large strains the arrangements are far more complex, and large variations between the walls are observed. It is a

general observation that the cell size and the misorientation across the cell walls depend on the plastic strain, and it is frequently observed that there is a decrease in cell size together with an increase in misorientation when the strain is in- creased [4,5,6J. Figure 2 illustrates the cell size-strain dependence for aluminium deformed by drawing, and it is noted that the cell size de-

creases continuously with increasing strain. This behaviour is typical when the deformation mode is drawing or rolling [4], whereas, for material de- formed in tension, the cell size may reach a con- stant value at a strain of the order of 0.1-0.2 [5,6]. An increase in strain above this values leads to an increase in the misorientation across

the cell walls indicating an increased dislocation density in the walls [6]. This effect of the defor- mation mode is at present unexplained and leads into the general problem as to how the dislocations

are arranged in the cell walls. Studies of such

Fig. 2. Cell size as a function of strain by cold drawing of 99.998% Al (from [2]). The material has been examined as-deformed and after a heat treatment at 140°C for 1 hour.

structures are quite difficult but modern microsco- pical and non-microscopical [7] techniques offer

new possibilities.

2.2. Bands.

The regular cell structure may contain a number of special features often in the shape of bands. One characteristic feature is ’dense dislocation walls’

(DDWs) [8]. These are structures running continu- ously and rather straight for distances of tens or a few hundreds of micrometers and aligned close to active slip planes. Their width and spacing are of the order of 0.2 pm and 2-4 pm, respectively. They

are narrow walls incorporating a greater misorien- tation than ordinary cells (up to 40). In figure 3 DDWs delineate the abrupt contrast changes between séquences of similarly oriented cells from lower left to upper right. The material in figure 3 is polycrystalline aluminium, but DDWs have also been observed in aluminium single crystals [9] as well

as in polycrystalline copper [10] and nickel [11].

On the basis of those findings it has been suggest- ed [12] that the DDWs are the counterpart of tilt walls parallel to the primary slip planes (mats) as observed in deformed single crystals.

Fig. 3. TEM micrograph of 15% cold-rolled Al - 0.6 vol.% A1203 (by courtesy of B. Bay). DDWs

delineate the abrupt contrast changes between

similarly oriented cells stretching from lower left to upper right. A microband is marked with

an arrow.

(4)

Fig. 4. TEM micrograph of 15% cold-rolled 99.998% Al showing the dislocation arrangement in two microbands (by courtesy of B. Bay).

Superimposed on the regular dislocation cell structure ’microbands’ composed of mueh smaller celles are often observed in metals such as alu- minium [13] and copper [14]. In general such bands contain a relatively high dislocation density, their width is a few tenths of a micrometer and they extend typically from tens to hundreds of micrometers. These bands are almost parallel to active slip planes, their spacing is typically 2-5 pm and the misorientation of the bands with respect to the surrcunding matrix is typically less than 2°. In Al-Mg far higher misorientations have been reported for local positions along microbands [15].

Examples of microbands in aluminium can be seen in figures 3 and 4. The microstructure of microbands in copper is more regular [141. Possible mechanisms for the formation of microbands and their role in the deformation process have been discussed in refs. [9,14,161 for instance.

The bands discussed above are related to the orien- tation of the slip systems in the grains in which they are found. In contrast, the orientation of shear bands does not correspond to the active slip systems and they are not limited by the grain boundaries. The characterization of shear bands and the mechanisms behind their formation will not be covered in the present paper.

In contrast to the bands described so far, other types of ’bands’ e.g. deformation bands, transition bands and grain boundary bands ineorporate larger misorientations across them (tens of degrees) and

extend over relatively large volumes of materials [4,17]. These bands start to form at relatively small strains and their number and the misorien- tation across them increases when the strain is increased. The occurrence of such bands reflects that grains can separate into regions deforming on

different slip systems or different combinations of slip systems or by various degrees of slip on a single set of slip systems. It appears that not only the lattice orientation but also the strain is subjected to local variations.

2.3. Surface observations.

Information complementary to that obtained in the studies of bulk dislocation structures can be ob- tained from slip line observations on the surface.

In pure f.c.c. metals slip line patterns include an elementary structure more or less uniformly distri-

buted over the surface consisting of fine lines with a spacing of the order of 20-40 nm which is normally only observed at low strains and a coarser structure of stronger slip lines which in medium and high stacking fault energy materials may cluster to form very coarse slip bands of variable spacing. The slip bands represent rather large shears (offset typically 1000 Burgers vectors or more). The distribution of slip lines/bands reflect the fragmentation described in 2.2. In the follow- ing some recent observations are presented.

A direct comparison of surface and bulk structures is possible by examining thin foils from a surface [8]. The structures obtained by this technique can be examined using a high voltage electron micro- scope. The results show that the dislocation structure reveals the direction of the active slip planes as indicated by the slip bands on the surface. Furthermore, DDWs and microbands appear to be individually associated with the corresponding slip bands. These observations indicate that the DDWs outline the large initial cells formed during primary slip and that slip in a cell structure

occurs over distances which may involve a number of cells. This latter suggestion is in accord with what is often observed in in situ straining experi- ments in a high voltage electron microscope, namely that dislocations stop at cell walls but then may continue after momentary arrest [18].

2.4. Effect of metallurgical parameters.

Metallurgical parameters as grain size, second phase particles, elements in solid solution etc.

affect the slip patterns and the resultant micro- structures. Most of these effects have not been

quantified, except for the cell size, which tends to decrease when solute elements and particles are introduced. Furthermore, the cell size is affected by the stacking fault energy of the base métal [5].

The effect of grain size and large and small par- ticles on the slip pattern will be discussed in the following.

a) Grain size. Surface relief patterns show that the slip line structure is finer in fine-grained specimens for small degrees of deformation. The slip pattern is inhomogeneous in nature. However, the non-uniform distribution of slip bands tends to be least pronounced in fine-grained specimens [19].

With increasing degree of deformation the slip pattern becomes quite similar in fine-grained and coarse-grained specimens [191.

Dislocation structures examined by TEM are inhomo-

geneous and the grain size appears to have rela-

tively little effect on the structures observed, but it is generally found that the dislocation density at a constant strain tends to increase when the grain size is decreased [17]. Grain fragmen- tation as studied by selected area channeling pat- terns appears to be less pronounced in fine-grained

(5)

than in coarse-grained spécimens (J.B. Bilde-Séren-

sen, private communication).

b) Large Darticles. The presence of coarse hard particles (> 0.1 pm) causes plastic inhomogeneity

at the interface between particles and matrix which, can lead to the formation of local deformation zones where the lattice is rotated significantly

with respect to the matrix [201. The orientation spread within the complète deformation zone has been determined by mierotexture measurements for single crystals [21] and it has been found that parts of the deformation zone at the particles have misorientations up to 40°, but a relative large fraction of the deformation zone has orientations very similar to that of the surrounding matrix.

Systematic studies have not been carried out on

polycrystalline materials but a number of findings

indicate that the effect of particles is similar to that observed for single crystals. However, it has also been-observed that rotations even at relative- ly large particles may be small [22]. Furthermore, it has been observed that particles agglomerated in groups and particles present in transition bands and at grain boundaries may lead to enhanced ro- tations and increased density of dislocations.

c) Small partioles. Small hard (non-shearable) par- ticles « 0.1 Nm) refine the slip pattern, and a general finding by SEM is that the slip lines are fainter, more closely spaced and often interrupted when compared with material without particles [231.

The dislocation structures in the bulk are also affected by the présence of particles, whieh leads to higher dislocation density and a smaller subgrain size at a given strain [2]. The formation

of structural features indicating lattice rotations such as transition bands and grain boundary bands

have been studied in Al-Al203 materials [23]. The

general observation is that the effect of particles is small except at low degrees of deformation, where the formation of transition bands is more pronounced in aluminium without particles. This finding seems to reflect a homogenizing effect of particles on the deformation pattern.

3. DEFORMATION TEXTURES

The deformation textures of f.c.c. materials have been studied extensively, particularly the rolling textures. Such studies combined with model calcu- lations can give important information about defor- mation mechanisms and about the effect of texture on the relationship between the flow stress and strain and on anisotropy. The emphasis in the pre- sent work will be on materials developing a copper- type rolling texture, in particular aluminium.

3.1. Rolling textures.

Figure 5 shows the texture of pure aluminium (99.998%) rolled 90 % as expressed in the {111}

pole figure. The aluminium texture is a copper- type texture (as opposed to the brass-type texture developed in materials with low stacking-fault energy). For comparison with figure 5 the {111}

pole figure for rolled copper is shown in figure 6.

These figures show that the textures in aluminium

Fig. 5. (1111 pole figure for pure aluminium rolled to 90% réduction (by courtesy of D. Juul Jensen).

RD

5 8

3 2 1

1

5 z

8 3

12 16

T D

1 2 3 5

RD 8

5

3 2 1

1

5 2

8 3

12 16

TD

1 2 3 5

Fig. 6. (1111 pole figure for OFHC copper rolled to 95% reduction.

and copper are comparable, although not identical.

The differences can be elucidated by comparing the experimental pole figures with those calculated on

the basis of texture models. The Taylor model [25]

and the relaxed-constraint model [26,27,28] produce the pole figures shown in figures 7 and 8 [291. A comparison of figures 5, 6, 7 and 8 shows that the texture of copper at high strains is very similar to the texture calculated by the relaxed-

constraint model, whereas the texture of aluminium is somewhere between the relaxed -constraint and the full-constraint texture. The same trend, i.e. that the texture is in between a relaxed-constraint and

a full-constraint texture, is also observed in the pole figures-for commercially pure aluminium quoted in the literature (e.g. [24]), but normally these pole figures come closer to relaxed constraint than that in fig. 5. It should be mentioned that textures very similar to the full-constraint texture in figure 7 have been observed experimen-

(6)

RD

TD Fig. 7. {111} pole figure for a simulated full- constraint Taylor texture (from [291). Contour lines for pole densities in the range 0.5-19.

Fig. 8. (1111 pole figure for a simulated re- laxed-constraint texture (from [291). Contour lines for pole densities in the range 1-13.

tally for spécial conditions, viz. rolling of alu- minium (and aluminium alloys) at elevated tempera- ture [24,30]. The texture of copper at lower

rolling réduction is, like the fully developed

aluminium texture, somewhere between relaxed con- straint and full constraint, but elosest to relaxed constraint [291.

3.2. Effect of metallurgical parameters.

Deformation textures are qualitatively and quanti- tatively affected by a number of metallurgical parameters including the stacking fault energy.

However, in the following we shall concentrate on the effect of grain size and dispersed particles (as in the section on microstructures) and hence neglect the problem of the formation of the brass-

type texture (e.g. [31,321).

a) Effect of 2rain size. The effect of grain size

on the textural development during cold rolling has

been studied in some detail [19,331, and it is found that an inerease in grain size reduces the rate of textural development and decreases the maximum orientation densities even at large degrees

of deformation (figure 9). In studying the effect of grain size, an inherent problem is that a vari- ation in this parameter normally is accompanied by

a change in the initial texture. Thus the textural development observed might be a combined effect of the two parameters. This has been taken into ac- count in an experimental investigation of the ef- fect of grain size on the texture and micro- structure of commercially pure aluminium (approx.

99.5%) cold rolled from 15 to 95% réduction in thickness at room temperature [34] and in an in- vestigation by computer simulation [29]. The results from these studies show that both the ini- tial grain size and the starting texture have pro- nounced effects on the development of the rolling texture and that these effects are significantly

reduced at large rolling réductions.

DEGREE OF DEFORMATION (%) Fig. 9. The development of rolling texture in fine-grained (50 pm) and coarse-grained (300 pm) aluminium (from [19]). The rolling texture is characterized by the sum of the orientations within 150 from (1101 112>, fl231 634> and [1121 111> in ODF space.

b) Effect of large Darticles. The effect of coarse hard particles (> 0.1 pm) on the deformation tex- ture has not been investigated systematically. How-

ever, a number of experimental findings for various materials indicates that the présence of large particles in general causes a weakening of the texture and that this effect becomes more pro- nounced when the volume concentration of particles is large [341.

The randomizing effect of large particles on the textural development can be related to the for-

(7)

mation of deformation zones at the particles where the lattice is rotated significantly with respect to the matrix. How much the deformation texture in

a polycrystalline material actually is weakened by the présence of large particles will therefore depend on the volume fraction of particles, the size of the deformation zones and the lattice ro-

tations within the zones. But besides that, the spread of the texture in the particle free matrix is important [341. This spread may depend on para- meters such as material, starting texture, grain size, degree of réduction etc. To determine the

weakening of the deformation texture the lattice rotations in the deformation zones around the par- ticles have to be compared with the spread of the matrix texture away from the skeleton line of the orientation tube whieh may be 5-10°, in some cases up to 15°. If the maximum lattice rotations in the deformation zones are small (5-15°) and therefore comparable to the spread of the matrix texture, most of the orientations in the deformation zones will be present in the matrix texture and the texture will only be slightly weakened. This would be the case in lightly deformed materials contain- ing relative small particles. However, at medium and large degrees of deformation the lattice rota- tions in the deformation zones can be much larger

than the spread of the matrix texture, e.g. 30-40°

and the texture will be significantly weakened [341.

c) Effect of small particles. The effect of small hard (non-shearable) particles « 0.1 Mm) on the

deformation textures have been investigated in a number of studies (see [34]), but it is difficult to establish a trend. For small volume concen- tration of particles some studies have shown that the texture is strengthened whereas in other cases a neutral or a weakening effect has been observed.

For large volume concentrations of particles, the général effect appears to be that the deformation texture is weakened.

One might imagine that homogenization of the slip pattern is the predominant effect for small volume fractions of particles and that this might produce texture strengthening. For large fractions the distortion of the slip pattern would dominate, resulting in a weakening of the texture.

4. DEFORMATION MODELS

In this section the existing models for the plastie deformation of polycrystals will be discussed in terms of the expérimental observations of micro- structure and texture reviewed in sections 2 and 3.

Only the observations on single-phase materials

will be dealt with. The knowledge of the effects of particles is so limited that any discussion of their model implications beyond the remarks already made in sections 2 and 3 would be premature. First the spectrum of existing models will be briefly

reviewed.

The Taylor model [25] describes a homogeneous slip pattern with basically no qualitative difference

between the different slip systems, only a quanti-

tative difference between the amount of shear. The relaxed-constraint model (e.g. [26,27]) relies on

heterogeneous plastic interaction between the grains to account for a minor part of material continuity, but the major fraction of slip is homo- geneous. As in the Taylor model there is only a quantitative difference between the different slip systems. The relaxation of constraints is linked to the shape of the grains at large strain (there is

no relaxation for equiaxed grains). Relaxation of constraint has been suggested earlier with dif- ferent arguments [281. Other models, like the ’mo- dified Sachs model’ (e.g. [35]), rely on plastic interaction between the grains (manifesting itself

in a heterogeneous slip pattern, e.g. [35,361) to account for a large part of’material continuity. In

the modified Sachs model there is a qualitative difference between the ’primary’ slip systems (the slip systems with the highest resolved shear stress from the applied stress) and the other systems.

Formally, the plastic interaction between the

grains is inversely related to the interaction with

a continuum matrix. In the Taylor model the indivi- dual grains have a very strong (rigid-plastie)

interaction with the continuum matrix and hence no

plastic interaction with the actual neighbours. The

more the formal interaction with the continuum matrix is relaxed (e.g. by relaxation of individual constraints), the more the models rely on plastic interaction between the grains for strain accom-

modation. The self-consistent scheme of Berveiller and Zaoui and co-workers (e.g. [37]) covers the whole range from (elastic-plastic) Taylor-type

deformation to deformation with strong plastic interaction by adjusting the strength of the inter- action with the continuum matrix. Even when plastic grain interaction is implied in a model, this does not necessarily mean that the corresponding slip is accounted for when the model is used for quanti- tative simulation (texture simulation for in- stance).

4.1. Models and microscopical observations.

In the Taylor model the number of active slip systems required and accounted for is 5 (on 3 or 4 slip planes in f.c.c. materials). In the relaxed- constraint model the number of slip systems formal- ly accounted for is marginally smaller, and in

models like the modified Sachs model it is substan- tially smaller. But when the slip systems not for- mally accounted for are added, the total number of slip systems would, in all models, come to at least 5 as in the Taylor model. Thus, the total number of slip systems is not the major issue in the micro- scopical investigation of the slip pattern. The important points are the homogeneity or hetero- geneity of the distribution of slip and the charac- ter of the different slip systems.

The slip pattern varies throughout the average

grain - as it seems more so in coarse-grained than in fine-grained materials. This is not basically incompatible with the Taylor model because of the ambiguity in the sélection of slip systems, but it does support models like the relaxed-constraint model or the modified Sachs model whieh positively

include heterogeneous deformation. Furthermore, the

Références

Documents relatifs

Then, fatigue tests on CT and oligocyclic specimen have been conducted on as-received and pre-deformed material to further develop a fatigue crack propagation model to predict

A simple kinetic model of the solar corona has been developed at BISA (Pier- rard & Lamy 2003) in order to study the temperatures of the solar ions in coronal regions

P. Centre de Recherche et Developpement, B.P. This model enables t o take into account simultaneously the crystallographic texture, the morphology of grains, the

So the highly densified ferrites with large grain can be obtained by the coprecipitated powder, addition of BaO.B,O, and sintering by the vacuum sintering

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

Nous montrons ici qu’une variante de la piste presque-périodique pour l’hypothèse de Riemann généralisée passe par une bonne connaissance des applications linéaires dont les

Pour toutes ces raisons, par souci d’équité, le Conseil tient à souligner la nécessité de considérer l’ensemble des circonstances actuelles : celles dans