You are on page 1of 23

Typical Textures in Sheet Metals

Chapter One

Introduction

1.1 Overview

In many alloy systems the interrelation between microstructure and properties contains the key information for
establishing the strategies for material development. Texture is an essential part of the microstructure and can play a
dominant role when the anisotropy of the material is large and crucial. Texture and anisotropy is a traditional field of
materials science that has continuously expanded for about 70 years. It will continue to grow at an increasing rate due
to the progress made in this area, particularly during the last two decades:

Experimental techniques for the rapid determination of texture on macroscopic and microscopic scales have
improved considerably and are widely available in many laboratories.

Methods of texture representation and texture analyses have been perfected.

Physical concepts for modeling texture evolution in laboratory tests and in industrial processes are developing
rapidly.

Measured texture data can be used, on a routine basis, as input for computer codes to evaluate and quantitatively
predict anisotropic material behavior.

This coincides with an enormous increase of manufactured materials in number and complexity, where at the same
time demands are strong to achieve optimal material behavior by the adequate design of microstructure. The actual
orientation distribution in a polycrystal is the result of the manufacturing process and thus texture contains detailed
information about the production history of a work piece. On the other hand, texture has a strong effect on properties
so that it contains easily accessible information on the interrelation between processing parameters and materials
performance. In this context, texture evaluation and application provides many highly important aspects for the
interrelation between microstructure and properties for process control and material performance and for identification
of the controlling mechanisms.

Two of the main uses of aluminum alloys are for beverage container bodies and automotive body panels. Enhanced
formability in aluminum sheet products used in stamping of automotive panels follows from control of texture through
control of the production processes. Quantitative texture analysis provides the essential feedback for tuning the
thermomechanical history on the basis of detailed insight into controlling mechanisms, thereby leading to desired
properties in the finished product. Economic savings through control of texture can be significant. Due to the large
production volume of sheet material, such as aluminum can body stock, great savings can be achieved through the
reduction of scrap causes by earing tear offs.

1.1.1 Crystallographic anisotropy of aluminum alloys

Crystallographic anisotropy of aluminum alloys is generally characterized by earing behavior which has been
extensively investigated either experimentally or theoretically. It has been found that earing is strongly determined by
crystallographic textures. The relation between earing and texture is thus of great interest both to academy and
industry. Minimizing the earing of products has always been the object of the can industry. The industry also demands
the prediction of product earing based on the original hot band earing. Thus a total understanding of earing behavior
during cold rolling is a prerequisite for a satisfactory prediction. A large amount of research work has been dedicated
to this purpose. It is well known that a recrystallization texture causes 90° earing and that a deformation texture results
in 45° earing. With increasing cold reduction, the intensity of the recrystallization texture of the hot band decreases
while that of the deformation texture increases. Correspondingly, the 90° earing of the hot band decreases with cold
rolling and eventually changes to 45° earing in the final product. It is also found that a strong recrystallization texture
of the hot band is good for a stable and minimal earing behavior during cold rolling.

http://archive.uky.edu/bitstream/10225/437/11chap1.htm[2011-02-07 오후 12:46:45]
Typical Textures in Sheet Metals

These qualitative results have been a great help in the control of thermomechanical processes and of product quality.
However, a more quantitative relation is highly desired and is required for automatic process control and on-line
monitoring of the formability of aluminum alloys. This is usually characterized as a relation between formability
parameters (r-value, percent earing) and orientation intensity/volume fraction or the ODF (Orientation Distribution
Function) coefficients, W lmn . In this paper this relation is expressed as the percent earing and the orientation intensity
of the texture components. In addition, influence of the rolling process has also been investigated and discussed.

1.1.2 Luders bands of aluminum alloys

The pressure on the automotive industry to produce lighter cars with reduced fuel consumption causes a demand for
new materials able to replace steel for car body panels. Some of the materials considered for this application are
aluminum alloys, especially Al-Mg alloys of the 5xxx series. Although they combine a reasonable strength with good
corrosion resistance and formability, their mechanical properties are still inferior to those of steel and further
improvement is thus required. Therefore a deeper understanding of the processes taking place during preheating, hot
and cold rolling and annealing of Al-Mg rolled sheets is of crucial importance.

One striking problem of Al- Mg alloys (5xxx) is the surface roughening caused by Luders bands. The appearance of
Luders bands is usually related to the Portevin -Le Chatelier (PLC) effect (serrated yielding) which is generally
ascribed to the effect of dynamic strain ageing ---the dynamic interaction between solute atoms and mobile
dislocations. A lot of work has been performed on the parameters which could influence the interaction of dislocations
and solute in alloys. These include the effects of strain rate, deformation temperature, solute concentration, solution
treatment and precipitation. Serrated yielding normally occurs after a critical strain, e c, and when the strain rate
sensitivity , reaches a negative value. The effect is found to increase with increasing Mg content. Quenched
samples after solution treatment are more likely to cause the PLC effect since more solute is kept in solid solution.
Processes that cause precipitation, which decrease the solute in solid solution, will decrease the PLC effect.

When the alloy and the deformation temperature or strain rate do not favor the formation of Luders bands or the
detection of the Portevin -Le Chatelier effect, there still exists at least three causes for flow localization: namely
thermal, textural or structural softening. It is easy to determine if local heating is responsible for the observed shear
banding. This occurs when large strain rates are associated with low thermal conductivity. By contrast, when adiabatic
heating may be reasonably neglected, it is not straightforward to distinguish textural from structural origins of flow
localization. The former, sometimes called geometrical softening, is related to crystallographic rotations which are able
to optimize the orientation of slip systems in the majority of grains. The latter process is illustrated by the pioneering
works on latent hardening performed on single crystals and is frequently invoked when a change of strain path is
involved. Structural softening is mainly related to the decrease of the strength of barriers to the glide of dislocations,
which in the case of change of strain path constitutes alien dislocations.

Because of the close association of crystal deformation mechanisms with surface roughening, the crystallographic
texture and local crystal lattice misorientations are expected to have an effect on roughness, and also on the PLC
effect. The texture effect on surface roughening or on the PLC effect has been little studied [1,2]. In aluminum alloys,
a notable difference in the Taylor factor, computed for a state of balanced biaxial tension, exists between grains of
distinct crystallographic bands [3,4]. Etching techniques reveal grooves that are populated with colonies of grains
having near-cube orientation [3]. The formation of a groove indicates that the cube-oriented colonies have (on the
average) lower through-thickness strength than the surrounding material. This supports the notion that mechanical
inhomogeneity arising from the spatial segregation of crystallographic texture can promote direction roughening,
which eventually leads to local failure. Experiments on copper sheets with a high cube texture suggest that surface
roughening may be reduced for materials with certain strong crystallographic texture [5]. This is because the
deformation incompatibility between neighboring grains is reduced. However, the cube orientation is not stable at
finite strains, and intense strain localization can development at a lower strain than for a randomly textured sheet.
Since the traditional x-ray techniques can only catch the texture information on a macroscopic scale (a statistical
meaning), they do not give the detailed orientation distribution of each grain, which is considered essential for the
investigation for the effect of texture on the PLC effect. With the Electron Back Scattering Pattern (EBSP), which first
appeared in the early 90's, it is now possible to obtain the orientation for each grain. Several works [1,6] have shown

http://archive.uky.edu/bitstream/10225/437/11chap1.htm[2011-02-07 오후 12:46:45]
Typical Textures in Sheet Metals

that the orientation distribution of grains is not random. The grains with similar orientations have a trend to get
together to form a "texture clustering". This phenomena has been found most recently both in SC and DC aluminum
alloys by Liu et al [7], and he has named it "texture continuity". Since the Luders bands are the product of
inhomogeneous deformation, which is closely related to the local lattice orientation, the orientation distribution of
grains is expected to have an effect on their formation. In this regard, the phenomenon of texture clustering implies a
possible effect on the PLC effect. This has been considered and is a totally different concept from the conventional
idea, and deserves more investigation.

1.1.3 Age softening behavior

Age softening of strain -hardened Al- Mg alloys is of particular commercial significance, because it leads to softening
of the heavily worked Al- Mg alloys when they are held for increasing times at room temperature. The age softening
effect increases with increasing cold work and with increase in Mg content. Since the PLC effect is also very strong in
Al-Mg alloys and increases in intensity with increase in Mg content it is reasonable to deduce that the PLC effect is
also involved in the age softening behavior. The PLC effect which induces Mg clusters around dislocations leads to an
increase in strength during cold working of these alloys. This increase in strength is greater than if the PLC effect is
absent. Thus, it could be assumed that a part of the age softening behavior is associated with the long time
decomposition of the Mg clusters that are bound to the dislocations. Strain energy would be released by this event and
a relaxation of the tangled dislocation structure would occur. It is recognized that Al-Mg solid solutions decompose
according to the reaction S.S.® b ''® b '® b and that this reaction is in part the driving force for the Mg clusters to
change with time in their characteristics and behavior.

It is to be expected that texture and anisotropy will gain increasing interest in all sorts of transformation processes
(such as recrystallization, precipitation and martensite formation). It is also assumed that the effect of interphases or of
local orientation arrangements on evolution of microstructure and texture will play an important role and should be
incorporated routinely in most of the investigations of the mechanical properties and formability during those
processes.

1.2 Texture measurements and representations

1.2.1 The representation of orientations and textures

The orientation of a crystal may be characterized in a great many different ways with respect to the sample axes. One
of the most frequently used representations of an orientation is to specify the crystal direction which is parallel to the
sheet normal direction (ND) and the one parallel to the rolling direction (RD)

g = (hkl)[uvw].

For the sake of symmetry it is now convenient to include the third perpendicular direction namely the transverse
direction (TD) and to specify the crystal direction in a matrix.

       RD   TD   ND

where the components are the Miller indices of RD, TD, ND. It is usually normalized to 1 as:

http://archive.uky.edu/bitstream/10225/437/11chap1.htm[2011-02-07 오후 12:46:45]
Typical Textures in Sheet Metals

where

For practical manipulations of orientations, the Euler angles have been introduced for these representations. By three
rotations of the axes, the sample frame is rotated to coincide with that of the crystal. There are two different rotations
given by Bunge [8] and Roe [9]. Here the Roe's method is adopted, which is shown as Figure1.

1. Rotate about z-axis by an angle Y.

2. Rotate about the new y'-axis by an angle Q.

3. Rotate about the new z''-axis by an angle F.

The orientation g is thus defined by

which is simplified as

Under the sample coordinates, the orientation g is represented by the matrix

Or

Comparing Eq.2 with Eq.3 one obtains the interrelation between the Euler angles and the direction cosines of the
rolling and normal directions and thus the relation between the Euler angles and the Miller indices.

1.2.2 Texture measurements

The preferred orientation of polycrystallities is usually determined by pole figure measurements. X-ray diffraction is
most commonly applied and will be discussed here. But other techniques, such as Neutron diffraction and Electron
diffraction using TEM or SEM, are also used for this purpose.

X-ray diffraction was first employed by Wever [10] to investigate the preferred orientation in metals. However, only

http://archive.uky.edu/bitstream/10225/437/11chap1.htm[2011-02-07 오후 12:46:45]
Typical Textures in Sheet Metals

with the introduction of the pole figure goniometer and the use of Geiger counters by Decker & al [11] and Norton
[12] has it become a quantitative method. Many aspects of pole figure goniometry have been discussed in some detail
by Schulz [13,14]. Bragg's law for monochromatic radiation is applied. It has two conditions. The first is that a lattice
plane (hkl) diffracts if it is in a reflection position between the incident and diffracted x-rays. The second condition is
that lattice planes with a spacing dhkl obey the law

where 2q is the angle between incident and diffracted beams, and l is the monochromatic x-ray wavelength and n an
integer defining the order of diffraction. The principle is simple (Figure2). In order to determine the orientation of a
given lattice, (hkl), of a single crystallite, the detector is first set at the proper Bragg angle, 2q, of the diffraction peak
of interest, then the sample is rotated in a goniometer until the lattice plane (hkl) is in the reflection condition (i.e. the
normal to the lattice plane or diffraction vector is the bisectrix between the incident and diffracted beam). The
goniometer rotations are related to the angular coordinates which define a sample orientation. In the case of a
polycrystalline sample, the intensity recorded at a certain sample orientation is proportional to the volume fraction of
crystallites with their lattice planes in reflection geometry.

Two methods of analysis are generally used: Determination of texture can be done on a sample of large thickness and
a plane surface on which x-rays are reflected (reflection) or on a thin slab of thickness t which is penetrated by x-rays
(transmission). Since x- rays are strongly absorbed by matter, the transmission method is only applicable on very thin
foils or wires (<100m m) and on materials with relatively low absorption. Among common metals, only Al, Mg, and
Be can be studied by the transmission technique, and most routine pole figure measurements are done using the
reflection method.

A pole figure collects a sum of lattice plane reflection signals from a large number of crystals. In the pole figure we
have not only lost spatial information, but also some orientations (such as how the axes of individual crystallites
correlate). It is not possible to determine the orientation density of crystallites in a polycrystal from densities in a pole
figure without a certain ambiguity. Therefore, unless the texture of a material is very simple, it is very difficult to
determine quantitatively from pole figures or inverse pole figures the amount of all the orientations present. Only a
three-dimensional plot of orientations can fully describe a texture. Such plots are often called orientation distribution
functions (ODFs) because they can be determined by mathematically representing the texture by spherical harmonics.

1.2.3 Determination of orientation distribution from pole figure

The most satisfactory way of obtaining explicit information from pole figure data is by using those data to compute the
three-dimension orientation distribution functions (ODFs), f(a ,b ,g ), of a sample. The relation between the ODF and
the Pole Figure is obvious from Figure2. The reflection condition is not changed if the crystal is rotated about the plane
normal (hkl) direction. Hence, the pole density is an integral over the orientation distribution function:

where 1/2p is just for the normalization purpose.

The computational algorithms for ODF analyses fall into two different categories, those in which the computations are
performed in Fourier space (harmonic methods), and those in which the computations are performed directly in
orientation space (direct or discrete methods). Here we talk about only the commonly used harmonic methods. The
basic premise of the harmonic method is that the pole figure and ODF are mathematically well balanced, and can
therefore be fitted by a series expansion with appropriate mathematical functions. Appropriate functions to use in a
spherical coordinate system are the spherical harmonic functions. The use of spherical harmonics to describe texture
was proposed by Pursey & Cox [15] and Viglin [16], and a complete scheme for texture analysis was proposed
independently by Bunge [8] and Roe [9]. Although their methods are conceptually identical, the formalism used by
Roe and Bunge is different. Here the Roe's formalism is used. The (hkl) pole figure, p(a , b ) is expanded in a series of

http://archive.uky.edu/bitstream/10225/437/11chap1.htm[2011-02-07 오후 12:46:45]
Typical Textures in Sheet Metals

spherical harmonic functions:

In this equation Qlm are coefficients to be determined, Plm (cosa ) e imb is a spherical harmonic function (P is an
associated Legendre polynomial), and l and m are integers that govern the shape of the function (these would be the
angular momentum and magnetic quantum numbers in the solution for the hydrogen atom); l is often called the 'order'
of the spherical harmonic function. This is an infinite series, but it is clear that truncation at some finite value of l is
necessary in practice, limited by the number of data points measured. This limits the resolution of the method and
leads to termination or truncation errors in the results. The coefficients are, in principle, complex, but symmetries in
the material often reduce the imaginary parts to zero, while the complex exponents in Eq. 5 can be reduced to cosine
functions.

Because the functions are orthogonal, the coefficients can easily be obtained from the experimental data P by
integration:

It is assumed that the ODF can be similarly expanded in a series of generalized spherical harmonic functions,
specifically:

where W lmn are the coefficients of this series, and Zlmn are Jacobi polynomials. The problem becomes one of finding
the relation between the unknown W lmn and the experimentally accessible. By substituting Eqs. 5 and 7 in Eq. 4,
integrating, and making use of the Legendre addition theorem, we obtain

Here x and h are the polar coordinates of the (hkl) pole in the crystal coordinate system.

Equation 8 is a linear equation relating the pole figure coefficients with the ODF coefficients. By measuring several
pole figures from geometrically independent poles, we obtain a set of linear simultaneous equations that can be solved
for the W lmn . The W lmn can then be substituted in Eq. 7 to obtain the ODF.

The number of pole figures that must be measured depends on the number of independent unknowns in Eq. 8. This in
turn depends on the order l to which we intend to compute the series. Symmetry in the crystal Laue group will reduce
the number of independent W lmn and the number of required pole figures. For example, cubic crystal symmetry
dictates that W 2mn =0 (and therefore Q2m=0), and in fact allows all the independent coefficients to be determined to l =
22 from only two pole figures [9].

Once the W lmn have been found Eqs. 8 and 5 can be used to re-calculate pole figures, even for pole figures that were
not measured or used in the analysis.

The simple calculation of pole figure coefficients, Q lm, from the experimental data using Eq. 6 is a consequence of the
orthogonality of the spherical harmonic functions over the surface of the sphere. With the available experimental

http://archive.uky.edu/bitstream/10225/437/11chap1.htm[2011-02-07 오후 12:46:45]
Typical Textures in Sheet Metals

methods, however, it is difficult in many cases to obtain complete pole figures, or at least, it requires a much larger
experimental effort than the determination of incomplete pole figure for example in the back -reflection range only.
Equation 6 cannot, therefore, be used as it stands with incomplete pole figures. Two ways have been used to overcome
this difficulty. One of these [17] uses a least-squares analysis to find the ODF coefficients that best fit the experimental
data.

A second method is much easier to implement, and involves an initial extrapolation from the known to the missing
part of the pole figures, usually with a third-order polynomial. For materials with cubic crystal symmetry the
requirement that Q2m=0 can be used to properly normalize the values in this extrapolation [18]. Equations 6 and 8 are
then used to calculate a first estimate of the W lmn . These values are used to recalculate the missing parts of the pole
figure (from Eqs. 8 and 5). Any physically meaningless negative values are replaced with zeros. These recalculated
parts contain information derived from, and crystallographically consistent with, the real experimental data. The
recalculated values are, therefore, much better estimates than the original extrapolated values, and are substituted for
them. The process is repeated, each iteration improving the reliability of the estimate of the missing parts, until the
self-consistency of the data is judged satisfactory. Dahms & Bunge [19] suggest that even better results will be
obtained by recalculating additional (non-measured) pole figures, correcting for any negative values, and then
incorporating these into the iterative process.

1.3 Typical textures in sheet metals

1.3.1 Deformation textures in fcc metals

The textures that have been most thoroughly investigated in metallurgy have been those in rolled sheets, or rolled and
recrystallized sheets, of materials of cubic lattice structure. Some textures are well described by 'components': a
superposition of a small number of single crystals, with some spread (which may be quantified by Gaussians) [20,21].
Others can be idealized as 'fiber' in orientation space, in which a single angle can be used to specify an orientation
within the fiber (although it may not be simply one of the Euler angles). Thus the analyzing of rolling textures has
usually been approached in a simplified manner: (1) using texture components and (2) using fibers. The aim of using
texture components is to reduce the representation of the orientation distribution into a small set of specific orientations
or texture components. Table 1 lists the Euler angles and Miller indices of the texture components that appear in most
fcc metals.

Table 1 Rolling texture components: Indices and Euler angles [22]

Name Indices Bunge Kocks

(j 1 ,F ,j 2 ) (y ,Q ,f )

Copper 90, 35, 45 0, 35, 45

S1 59, 29, 63 149, 29, 27

S2 47, 37, 63 137, 37, 27

S3* 59, 37, 63 149, 37, 27

Brass 35, 45, 0 55, 45, 0

Taylor {4 4 11}<11 11 90, 27, 45 0, 27, 45


>

http://archive.uky.edu/bitstream/10225/437/11chap1.htm[2011-02-07 오후 12:46:45]
Typical Textures in Sheet Metals

Goss 0, 45, 0 90, 45, 0

* This particular orientation is often quoted as 'the' characteristic S orientation. Nevertheless, there is significant
variation in the literature as to which precise orientations are labeled as 'S'.

The other approach for studying rolling textures uses the concept of partial fibers, where a fiber is a range of
orientations limited to a single degree of (rotational) freedom about a fixed axis [23]; some fibers are defined in terms
of an axis in sample space, and some in terms of crystal coordinates. The appearance of a fiber in ODF plots is that of
a line, which may or may not lie entirely in one section. There are two fibers typically used for fcc metals: the b and a
fibers. The b fiber runs (in orientation space) from the 'copper' to the 'brass' component. The 'S' orientations are at
intermediate locations along this fiber. The a fiber runs from 'brass' to 'Goss'. The b fiber is generally prominent in
rolling textures. Ideally, these fibers should always correspond to regions of high intensity in the orientation
distribution. However, the location of the maximal values in any given ODF may vary significantly from the ideal.
This has led to definitions of fibers that are not fixed in position ('skeleton lines') [17].

It is often assumed that the texture of a polycrystal rolled to a high reduction will give essentially the same texture as a
polycrystal rolled to a lower reduction only stronger. While a higher reduction will generally produce a sharper
texture, the character of the texture may change significantly. )LJ3 shows (111) pole figures for copper rolled to
increasing reductions [23]. In this figure, the texture sharpens with increasing strain, as would be expected, and the
general features of the texture remain relatively constant as well. However, the locations of the peaks in the pole
figures shift considerably in orientation space. As discussed below, this means that caution is required when analyzing
textures in terms of fibers that are fixed in orientation.

The variation in rolling texture with material has been well documented by previous researchers. It is convenient to
regard the 'copper' texture and the 'brass' texture as the opposite extremes of 'pure metal' and 'alloy' textures (see Table
1 for definitions). The effect of increasing the amount of zinc in copper is shown in )LJ4[24]. This set of (111) pole
figures shows the transition from the pure metal to the alloy type texture over a range of Zn content in which twinning
is not generally observed. The figure shows that there is a transition to the alloy texture at about 15% zinc. However,
this transition can also be achieved with the addition of only a few percent of phosphorus [25].

Similarly, elevated deformation temperatures tend to lead to the pure metal texture, whereas low temperatures have the
opposite effect. For example, rolling copper at liquid nitrogen temperature yields a brass texture [26], and rolling silver
at high temperature shifts the texture toward that of Al [27,28]. Figure5 [29] shows the effect of temperature on texture
evolution in 95% rolled plutonium with 9.6 atomic percent gallium. These pole figures show that the transition from
the pure metal to the alloy type texture can be achieved by decreasing the deformation temperature. A rather complete
table is available in Dillamore & Roberts [25] summarizing the texture data for most fcc alloy systems.

It is also worth noting the controversy in the literature concerning the role of microstructure on the formation of the
alloy or brass texture. Although brass itself twins readily at small strains, it is not clear that additional twinning occurs
at higher strains (greater reductions in thickness) during rolling at which the brass component becomes strong. In fact,
the brass component appears without twinning in many cases; for example, this component is often dominant in
aluminum-lithium alloys [30], obviously for different reasons.

1.3.2 Recrystallization texture in rolled fcc metals

The topic of recrystallization texture is not merely of academic interest since the properties of many commercial alloys
depend critically on the control of the recrystallization texture. Examples abound in the literature on silicon-bearing
steels, used for their magnetic properties, and aluminum sheet for beverage cans. On the other hand, recrystallization
textures in rolled copper have been investigated most thoroughly from a fundamental point of view. Copper is more
easily obtained in high purity whereas aluminum very often has significant levels of both insoluble and soluble
impurities.

As an illustration of the wide variety of recrystallization textures that can be obtained in fcc materials, Figure6 shows

http://archive.uky.edu/bitstream/10225/437/11chap1.htm[2011-02-07 오후 12:46:45]
Typical Textures in Sheet Metals

the (111) pole figures of the recrystallization textures of Al, Cu, Brass, and an aluminum alloy. Recrystallization
textures (just like rolling textures) show significant material dependence, especially the effect of stacking-fault energy
(SFE). In materials of relatively high SFE (Al, Ni, Cu), a cube component often plays a major role. In addition to the
cube component, Al and Al alloys often exhibit a component close to the typical rolling texture S-orientation
(Fig.6a)[31]. In Ni and Cu, recrystallization twinning leads to minor intensities of the cube orientation's first-
generation twin. ( Fig.6b). In contrast to these examples, low-SFE materials (Ag, brass, austenitic steels) generally
exhibit much weaker recrystallization textures, which are characterized by the so-called brass recrystallization
orientation {236}<383>(Fig.6c).

In comparison to rolling textures, recrystallization textures are much more complicated. In addition to the dependence
on SFE, the following conditions of the deformed state, the annealing temperature, and particularly the precipitation
state are known to exert strong influences on the recrystallization behavior and, consequently, on the recrystallization
textures [32]. As an example, the recrystallization texture of an Al-Mn alloy (AA3104), which was pretreated to
contain large (>1m m) constituent particles, is shown in Fig.6d. In comparison to pure Al (Fig.6a), the texture of the
two-phase material is much less pronounced, which is caused by the additional, nearly random nucleation in the
deformation zones around the second-phase particles ('particle stimulated nucleation') [33].

One overriding feature of recrystallization textures that has yet to be quantitatively understood is the development of
high symmetry components, such as cube ({100}<001>) and Goss ({110}<001>), after rolling [34]. As an example of
recent research on recrystallization in copper [23], Figure7 shows the evolution of the texture from the standard fcc
rolling texture, after 90% reduction in thickness by rolling, to a fully recrystallized condition. The initial texture of the
oxygen-free-electrical conductivity (OFE) copper before rolling had a weak cube component. The final texture after
rolling and annealing is an extremely strong cube texture with a very small volume fraction of a twin-related derivative
of the cube. This evolution is of considerable practical interest to the suppliers of aluminum sheet to the beverage
industry because the balance between deformation textures and annealing textures is crucial to controlling earing in
beverage can stock. It happens that the anisotropy of interest, most easily characterized in terms of the variation of R-
value with direction in the rolling plane has the same symmetry (four ears), but is turned by 45° , for the rolling and the
cube textures. Thus, it can happen that the texture in the partially recrystallized state (Fig.7b-e) is far from random, yet
the anisotropy of interest is nearly absent. This point is a caution against using any particular test of mechanical
isotropy as a diagnostic test for texture.

A more detailed analysis of the evolution of texture during recrystallization is shown in Figure8. This figure shows the
variation in volume fraction with fraction recrystallized for both the deformation and the recrystallization components.
The first thing to note in this figure is that the volume fractions of the deformation components decrease uniformly.
This indicates that no particular rolling component is depleted faster than any other during recrystallization. The steady
decrease in the deformation components is in contrast to the non-linear increase observed for the cube component. The
volume fractions in this case were calculated by first representing the orientation distribution by sets of weighted
discrete orientations [35]. The volume fraction for a specific component was then calculated by adding up the weights
of all of the weighted discrete orientations which were within 7.5° of the particular component and dividing this sum
by the weights of all of the discrete orientations in the set. (These results compared closely with volume fractions
calculated by integrating the intensity of the orientation distribution over a 15° ´ 15° ´ 15° volume around each
component.).

Another interesting aspect of texture evolution during recrystallization is the dependence on the deformed state.
Figure9 shows the variation in the cube component volume fraction with prior rolling strain, expressed here in terms of
the von Mises equivalent strain. Note that four different criteria have been used for the spread in the cube component
and that, at high pre -strains, most of the cube component oriented material is within 7.5° of the ideal position. The
results indicate that there is a marked transition from very little cube at pre-strains less than 2.0, to cube-dominated at
strains over about 3.0. Many other scalar measures of texture are available for these sorts of investigations, but if
textures are to be compared with microscopy then volume fractions of components are more appropriate than
intensities.

Lastly grain growth can also lead to changes in texture because of variations of grain boundary energy and mobility

http://archive.uky.edu/bitstream/10225/437/11chap1.htm[2011-02-07 오후 12:46:45]
Typical Textures in Sheet Metals

associated with specific texture components [36].

1.4 Anisotropy and formability of aluminum alloys

Enhanced formability in aluminum sheet products used in stamping of automotive panels follows from control of
texture through control of the production processes. Quantitative texture analysis provides the essential feedback for
tuning the thermomechanical history on the basis of detailed insight into controlling mechanisms, thereby leading to
desired properties in the finished product. Economic savings through control of texture can be significant. Due to the
large production volume of sheet material, such as aluminum can body stock, great savings can be achieved through
the reduction of scrap caused by earing.

1.4.1 Anisotropy behavior of aluminum sheets for can body

One of the main uses of aluminum alloys is for beverage containers. During the deep drawing process, undulations on
the rim of the formed article may be found. The high points are called "ears" and the general phenomenon is called
"earing". Minimizing earing is an important objective when producing rolled alloys that have to be subsequently deep
drawn. A randomly oriented material would be ideal for this purpose, since it would not show earing. This can be
achieved through multidirectional rolling or by small rolling reductions and frequent intermediate heat treatments, but
this is neither a practical nor an economical proposition.

Table 2. The effect of processing variables on earing of finally annealed materials

Fabrication variable Comments on effect of variable

Method of casting and casting Increased solidification rate tends to produce


conditions (DC vs. SC) more 45° earing and increase the range. Both
initial texture of cast product and ingot geometry
will influence the subsequent fabrication
practice.

Composition of material In dilute alloys decreasing (increasing Fe


content) purity produces more 45° earing; lower
Fe/Si content reduces the range. Minor additions
can influence earing in either direction
dependent on the Fe and Si contents.

Homogenization and Precipitation of supersaturated elements


preheating increases 45 ° earing. In commercial purity
materials 45 ° earing increases with increasing
temperature up to ~570° C.

Hot rolling Conditions that give rise to recrystallization


increase 90° earing.

Interannealing (including hot Normally increases 90° earing. The effect is


mill slab) reduced on repeated interannealing.

Cold rolling Increasing heavy reductions promote 45° earing.


After interannealing, results depend on the
degree of subsequent cold reduction.

Table 2 [37] summarizes some of the main factors in processing sheet aluminum, which can influence the earing of the

http://archive.uky.edu/bitstream/10225/437/11chap1.htm[2011-02-07 오후 12:46:45]
Typical Textures in Sheet Metals

final product. Because the fabrication of sheet materials involves a series of processes (casting, homogenization,
hot/cold rolling, annealing, etc.), it follows that almost every change in practice can potentially influence the ultimate
earing behavior of aluminum alloys. Therefore, the earing of aluminum alloys products can be controlled, to a great
extent, by the chemical composition, the casting process, ingot/strip homogenization, hot and cold rolling practices and
final annealing operations. Table 2 is not comprehensive and there are many more factors which have been observed to
influence earing behavior [38,39]. It is generally accepted that earing results from the presence of preferred orientation
in the metal being fabricated.

A number of factors influence the earing behavior of aluminum alloys by affecting directly or indirectly the preferred
orientation of the materials. In aluminum alloys, the earing has two main forms: (1) 45° earing, which is associated
with deformation type textures, and (2) 90° earing, which is associated with recrystallization type textures. The control
of earing in aluminum alloys depends therefore in a large measure on the balance between these two types of textures
in order to produce "earing -free" material. The combined effect of deformation and recrystallization on the
microstructure of a work piece is due to a complex interplay between the different physical processes of deformation
and recrystallization, which is not easily treated theoretically but, nevertheless, offer a wide range of possibilities for
manipulation of texture. For metallurgical applications an enormous amount of empirical knowledge has been gathered
and has led to standardized treatment schemes. In practice, when the hot band is to be cold rolled, the texture and
microstructure have essentially already been determined. It is in this sense that cold rolling determines the property of
the final products. Understanding the earing behavior during cold rolling is not only essential for predicting the
resultant formability, but also gives an insight into the thermomechanical processes needed to obtain a suitable original
hot band condition.

Earing behavior of aluminum alloys has been extensively investigated either experimentally or theoretically [38,40-
45]. It has been found that earing is strongly determined by crystallographic textures. The quantitative relation between
earing and texture is thus of much interest both to academy and industry. Minimizing the earing of products has
always been the objective of the can industry. The industry also demands the prediction of product earing based on the
original hot band earing. Thus a total understanding of earing behavior during cold rolling is a prerequisite for a
satisfactory prediction. A large amount of research work has been dedicated to this purpose [46,47]. It is well known
that the recrystallization textures cause 90 ° earing and that the deformation textures result in 45 ° earing. With
increasing cold reduction, the intensity of the recrystallization textures of hot band decreases while that of the
deformation textures increases. Correspondingly, 90° earing of the hot band decreases with cold rolling and eventually
changes to 45° earing in the final product. It is also found that a strong recrystallization texture of the hot band is
desirable for a stable earing behavior during cold rolling [46]. The can industry considers that 0/90° earing is more
harmful in the final sheet material and more easily causes an unacceptable number of defects than does 45° earing.
Therefore an investigation of the evolution of crystallographic textures of aluminum alloys and their earing is of
significant importance both to the can industry and for texture theory.

1.4.2 The Portevin-Le Chatelier (PLC) effect of aluminum alloys

In order to increase the ratio of strength/weight, replacing most of the steel with aluminum alloys is the current trend
for the automotive industry. Al-Mg alloys (5xxx) are most suitable for this purpose and currently are extensively
investigated. The striking problem is the surface quality caused by Luders bands. Luders bands forming in aluminum
alloys has been known for many years. The 5xxx series alloys (Al-Mg) of Al are most prone to this behavior. With the
appearance of Luders bands, the materials present a serrated yielding in the strain-stress curve which is designated the
Portevin-Le Chatelier (PLC) effect.

The occurrence of the Portevin -Le Chatelier effect (serrated yielding) is generally ascribed to the effect of dynamic
strain ageing ---the dynamic interaction between solute atoms and mobile dislocations. It normally occurs after a
critical strain e c and when the strain rate sensitivity D s / D ln reaches a negative value. Former work has shown the
existence of three types of serrated yielding, characterized by both the form of serrations, as evidenced on a stress-
strain curve, and the strain rate or temperature dependence of the critical strain, e c. At low and intermediate
temperatures or high strain rates in the serration range, e c is found to increase with increasing strain and decreasing

http://archive.uky.edu/bitstream/10225/437/11chap1.htm[2011-02-07 오후 12:46:45]
Typical Textures in Sheet Metals

temperature. In this region the solute mobility, which is initially insufficient to form atmospheres, is increased by
vacancy production during deformation while the dislocation velocity decreases, allowing the locking condition to be
met at e c. At e c the dislocations are effectively locked and further deformation requires dislocation breakaway and
multiplication, as evidenced on a stress -strain curve by an abrupt rise in flow stress followed by a discontinuous fall
back to or below the level of the curve [48-50]. Such serrations are called locking serrations as they originate from the
locking of initially mobile dislocations. In addition to vacancy concentrations produced by deformation, quenched-in
vacancy supersaturations have also been found to influence the strain at the onset of serrated flow [51].

At high temperatures and low strain rates, e c is often found to increase with decreasing strain rate and increasing
temperature [51-58]. In this region the form of the serrations is found to differ from the locking serrations in that, at
least initially, no increase in flow stress is observed prior to the discontinuous yield drop. Although little work has
been done in this region, it has been shown that the dislocation velocity is low enough initially for the dislocations to
be aged from the start of deformation. The serrations appear to be due to the breakaway and multiplication of initially
aged dislocations and are termed unlocking serrations. Thomas [58] has shown that each unlocking serration
corresponds to the initiation of a local Luders band.

Some experimental results have revealed a decreasing e c with increasing temperature [59 -62], whereas some have
also shown that e c decreases initially to a minimum then increases with temperature [63-65]. Regardless of the
difference in critical strain behavior, there exists a temperature range in which e c decreases with increasing
temperature. In this temperature range, the serration stress amplitude, D s , increases with increasing strain and
temperature [59,61,62] and decreases with increasing grain size [61,66,67]. Grain size plays a major role in
determining the strain to the onset of serrated yielding in a polycrystalline aluminum-3% magnesium alloy [52].

Various models [68-73] for dynamic strain ageing have been developed to account for the occurrence of serrated flow
at temperatures where solute diffusivity is extremely low, and for the observations of a critical strain and its
dependence on strain rate and temperature T and of the dependence of strain rate sensitivity of flow stress on strain
or stress. The vacancy model [69,70] considers the enhancement of the diffusion coefficient by the vacancies generated
during plastic flow [68], the increase in mobile dislocation density with strain [52] and the ageing of dislocations
during their waiting time at discrete obstacles [69,74]. In addition, the critical strain is interpreted in terms of the strain
at which the strain rate sensitivity becomes zero [70]. Negative strain rate sensitivity as a crucial factor for the onset of
serrated flow is now recognized in all the models.

In the models of Van den Beukel [70] and Mulford and Kocks [71] it is held that any solute mobility makes a negative
contribution to the total strain rate sensitivity and that this negative contribution increases with strain; when the total
strain rate sensitivity becomes negative, plastic flow becomes unstable. In arriving at the strain dependence of the
strain rate sensitivity Van den Beukel [70] retains the notions of a strain -dependent solute diffusion coefficient and
strain dependent mobile dislocation density from the older models. Mulford and Kocks [71], on the other hand,
describe the strain rate sensitivity in terms of the flow stress s which is the sum of two components---the friction stress
s f and the dislocation flow stress s d ; solute mobility affects s d , the strain hardening component of the flow stress.
They contend that the rate sensitivity of s d is negative, which in turn causes serrated flow. This interpretation of
critical strain does not rely on the production of vacancies for explaining the onset of serrated flow, instead it assumes
the diffusion of solute atoms along dislocations at forest intersections. In a joint paper, Van den Beukel and Kocks
[72] have arrived at a unified approach which enables the negative strain rate sensitivity to arise as a consequence of
the influence of solute mobility on both s f (by decreasing the obstacle spacing along the dislocation) and s d (by
increasing the strength of dislocation junctions). However, it should be mentioned that the vacancy concentration does
not significantly affect the initiation of jerky flow [54,75 -77]. In some models [71,73,74,78-80] the critical strain for
the occurrence of jerky flow may be explained without any assumption that vacancies created during deformation
contribute to the diffusion process. More recently, Kubin and Estrin [73] have proposed a model explaining the critical
strains for the appearance of plastic instability in terms of the strain dependence of the densities of mobile and forest
dislocations. In their model, the explanation of experimentally observed strain ranges of the PLC effect is not
connected directly to the particular choice of a model for dynamic strain ageing. Though there are objections raised

http://archive.uky.edu/bitstream/10225/437/11chap1.htm[2011-02-07 오후 12:46:45]
Typical Textures in Sheet Metals

against the models which rely on the vacancies generated during plastic flow, the vacancy model cannot be rejected, as
this would cause serious problems in explaining the differences in observed e c vs variation between interstitial and
substitutional solid solutions.

The various investigations on the serrated flow suggest that the measurement of e c and its dependence on and T is
essential in order to understand the underlying mechanisms. This dependence is generally expressed [69,81] as

(9)

where m and b are the respective exponents in the relations for the variation of vacancy concentration Cv and of
mobile dislocation density r m with plastic strain (Cvµ e m ; r µ e b ), K' is a constant, Q is the activation energy, k and
T have their usual meanings. One can obtain the exponent (m+b ) as the slope in the plot of ln vs lne c at a constant
temperature.

Because of the close association of crystal deformation mechanisms with surface roughening, the crystallographic
texture and local crystal lattice misorientations are expected to have an effect on roughness. Experiments on copper
sheets with a high cube texture suggest that surface roughening may be reduced for materials with certain strong
crystallographic textures [5]. This is because the deformation incompatibility between neighboring grains is reduced.
However, the cube orientation is not stable at finite strains and intense strain localization can develop at lower strains
than for a randomly textured sheet.

Another factor known to interact with surface roughening is small -scale strain localization near the surface [82,83].
Formation of bands of localized deformation can enhance roughening by increasing the depths of the valleys and
altering the surface strain distribution. Models formulated to capture shear bands near a free surface have assumed the
existence of a vertex at the loading point of the yield surface [84] or that the bands are related to the presence of
porosity [85].

1.5 Proposed work

Based on the aforementioned literature work this paper does not intend to study the formability of aluminum alloys per
se. It will, however, investigate different factors which affect the formability of aluminum alloys basically Al-Mg
alloys. Therefore, the following work will be carried out.

1.5.1 Texture and anisotropy of aluminum alloys

The crystallographic texture of aluminum alloys has been thoroughly investigated for different thermomechanical
processes. The anisotropy and formability of aluminum alloys have also been extensively studied either for their
dependence on crystallographic texture or for their direct dependence on different thermomechanical processes. The
qualitative relation between texture and anisotropy has been well developed. However, the need for an automated on-
line control of texture and formability is highly desired by industry and this requires a more quantitative relation
between texture and formability for certain processes. This is usually characterized as a relation between formability
parameters (r -value, percent earing) and orientation intensity/volume fraction [46,38] or the ODF coefficients, W lmn
[39,86-88]. In this work this relation is expressed as the percent earing and the orientation intensity of the texture.

For the can industry, when the hot band is ready to be cold rolled its microstructure and texture have already been
determined. In addition the final product formability is significantly determined by the cold rolling process. So a better
understanding of the earing behavior and texture evolution during cold rolling is very important for on- line texture
control. Thus, we want to perform work on the earing behavior and texture evolution during cold rolling. The main
purpose is to try to find an analytical relation between texture and earing. In addition, influence of the rolling process
was also investigated and discussed.

http://archive.uky.edu/bitstream/10225/437/11chap1.htm[2011-02-07 오후 12:46:45]
Typical Textures in Sheet Metals

1.5.2 PLC effect on Al-Mg alloys

The PLC effect on aluminum alloys has been extensively studied by the appearance of yielding serrations, i.e., the
intensity of the serrations, the onset of serrations, types of serrations, the magnitude of the stress drop as well as the
frequency. Many factors, i.e., temperature, strain rate, heat treatment process, composition, have also been studied with
regard to their effect on these parameters. Most of those investigations are based on the theory of dynamic strain
ageing---dynamic interaction between solute atoms and mobile dislocations first presented by Cottrell. However, when
the alloy and the deformation temperature or strain rate do not favor the formation of Luders bands or the Portevin-Le
Chatelier effect, there still exists at least three causes for flow localization: namely thermal, textural or structure
softening. It is rather easy to determine if local heating is responsible for the observed shear banding. This occurs only
if large strain rates are associated with low thermal conductivity. By contrast, when adiabatic heating may be
reasonably neglected, it is not straightforward to distinguish textural from structural origins of flow localization. The
former, sometimes called geometrical softening, is related to crystallographic rotations which are able to optimize the
orientation of slip systems in the majority of grains [89]. The latter process is illustrated by the pioneering work on
latent hardening performed on single crystals [90] and is frequently invoked when a change of strain path is involved
[91-93]. Structural softening is mainly related to the decrease of the strength of the barriers to the glide of dislocations,
which in the case of change of strain path constitutes alien dislocations. Most recently, Lopes et. al have shown the
textural instability in pure aluminum [94]. R. Becker has also shown that grain structure and texture has an effect on
the surface roughening during sheet forming [2].

It is proposed that this study of the PLC effect be carried out in relation to the age softening behavior and the micro
inhomogeneity of texture on Al-Mg (5xxx) alloys by chemical composition variations (alloy effect) as well as by
thermomechanical process variations (homogenization, solution treatment, annealing, deformation, etc.). This work
will incorporate both Direct Chill Cast (DC) material as well as Strip Cast (SC) material.

Copyright 2000 Ó Xiang-Ming Cheng

http://archive.uky.edu/bitstream/10225/437/11chap1.htm[2011-02-07 오후 12:46:45]

You might also like