You are on page 1of 26

Available online at www.sciencedirect.

com

Acta Materialia 61 (2013) 818843


www.elsevier.com/locate/actamat

Superior light metals by texture engineering: Optimized aluminum


and magnesium alloys for automotive applications
J. Hirsch a,, T. Al-Samman b,
b

a
Hydro Aluminum Rolled Products GmbH R&D, 53117 Bonn, Germany
Institut fur Metallkunde und Metallphysik, RWTH Aachen, 52056 Aachen, Germany

Abstract
Aluminum and magnesium are two highly important lightweight metals used in automotive applications to reduce vehicle weight.
Crystallographic texture engineering through a combination of intelligent processing and alloying is a powerful and eective tool to
obtain superior aluminum and magnesium alloys with optimized strength and ductility for automotive applications. In the present article
the basic mechanisms of texture formation of aluminum and magnesium alloys during wrought processing are described and the major
aspects and dierences in deformation and recrystallization mechanisms are discussed. In addition to the crystal structure, the resulting
properties can vary signicantly, depending on the alloy composition and processing conditions, which can cause drastic texture and
microstructure changes. The elementary mechanisms of plastic deformation and recrystallization comprising nucleation and growth
and their orientation dependence, either within the homogeneously formed microstructure or due to inhomogeneous deformation, are
described along with their impact on texture formation, and the resulting forming behavior. The typical face-centered cubic and hexagonal close-packed rolling and recrystallization textures, and related mechanical anisotropy and forming conditions are analyzed and
compared for standard aluminum and magnesium alloys. New aspects for their modication and advanced strategies of alloy design
and microstructure to improve material properties are derived.
2012 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Keywords: Aluminum; Magnesium; Texture; Deformation; Recrystallization

1. Introduction
The use of light structural materials has become inevitable in the modern world. The increasing need for improved
fuel economy has created a huge interest in lightweight automotive structures. Automotive structures can be best lightened through innovative design strategies directed toward
weight saving (e.g. thin-walled components) and by employing lightweight materials, such as aluminum and magnesium, which have the lightest density of all common
structural materials (qAl = 2.7 g cm3, qMg = 1.7 g cm3).
These two metals are positioned close to each other in the
periodic table and are quite similar in a number of basic
Corresponding authors.

E-mail addresses: Juergen.Hirsch@hydro.com (J. Hirsch), alsamman@imm.rwth-aachen.de (T. Al-Samman).

properties, such as atomic weight, elasticity, strength and


melting point. Both metals have close-packed atomic lattices
but dier in their possible variants, i.e. face-centered cubic
(fcc) and hexagonal close-packed (hcp), which explains their
fundamental dierences in forming behavior and singlecrystal plastic anisotropy.
While aluminum has already established a leading role
in automotive applications, the use of magnesium for parts
of automobile structures is still rather limited. Magnesium
shows limited ambient temperature formability due to a
shortage of independent deformation modes that would
accommodate deformation along the c-axis of the magnesium hcp unit cell, whereas aluminum with its four slip
planes, each with three slip directions, has 12 slip systems,
and thus, unconstrained formability. To satisfy a general
strain tensor, magnesium has to add additional but not
readily available glide systems and/or mechanical twinning,

1359-6454/$36.00 2012 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.actamat.2012.10.044

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843

which makes it prone to orientation eects caused by its


pronounced texture development during thermomechanical processing.
Another major dierence between the two materials is
the texture variation during recrystallization of the
deformed structure. In aluminum, depending on alloy composition and processing conditions, dramatic texture
changes are observed, due to strong selective recrystallization mechanisms during nucleation and nucleus growth.
Such dramatic but characteristic orientation changes are
missing in conventional magnesium and its alloys, where in
most cases the deformation texture is preserved during
annealing. These texture eects can be best analyzed in
cold-rolled (i.e. plane strain deformed) and annealed
sheet, where stress and strain conditions are well dened
and characteristic textures develop. For aluminum, typical
fcc rolling and recrystallization textures occur that aect
sheet formability mainly by the related anisotropy of
strength, ductility and related phenomena (yield locus, forming limit, earing). In magnesium a strong basal texture forms
rapidly during (hot) rolling and remains virtually unchanged
during recrystallization annealing, which aects subsequent
sheet formability. To improve magnesium sheet formability
one needs to depart from standard basal textures.
The following highlights the role of texture engineering
by intelligent processing and alloying as an eective tool
for optimizing aluminum and magnesium alloys for automotive applications. The texture evolution and its impact
on properties will be analyzed and compared for aluminum
alloys (i.e. AA5xxx AlMgMn alloys and AA6xxx Al
MgSi alloys) and most common magnesium alloys
(containing Al and Zn) that are predominantly used for
automotive lightweight engineering. Special attention is
paid to recent magnesium alloy development through microalloying with rare earth elements. The basic mechanisms
involved in deformation and recrystallization are analyzed
and compared. The role of alloy additions, solute segregation and second-phase precipitates, initial textures, deformation and annealing conditions, microstructure
heterogeneities and related recrystallization mechanisms
are considered, and possibilities are derived for their modication to guide potential strategies of alloy design and
microstructure engineering for optimal material properties.
2. Magnesium alloys
The tendency of magnesium and its alloys to develop
strong single-ber textures during thermomechanical processing creates major problems for forming at ambient
temperature. The key features of ductile magnesium alloys
are a homogeneous microstructure free of brittle intermetallic particles, and both a grain size and crystallographic
texture that promote uniform plastic deformation. Thus
to begin with, we will review the recent progress in obtaining ductile magnesium alloys via texture optimization that
serves to delay failure by permitting stable plastic deformation under low stresses. Particular focus is placed on the

819

favorable eects that rare earth alloying addition can have


on texture sharpness and in generating non-conventional
rare earth textures during conventional deformation
processes.
2.1. Advantages and limitations of magnesium alloys
Since the 1990s the forgotten material magnesium and
its alloys have been receiving widespread attention as playing an increasingly important role in almost every sector of
the metal consuming industry. In particular, the past decade has seen rising demand for magnesium alloy development for structural and automotive applications due to
their large potential for weight saving and, thereby, for
improved fuel economy and decreased exhaust emissions.
The range of applications of magnesium alloys has been
based on liquid and semi-solid molding processes and
sheet-forming processes. The vast majority of all magnesium components (automobile, housings and electric) are
produced by die-casting. Products take advantage of the
numerous appealing properties of magnesium, such as
lightness, high specic strength, good machinability, high
damping capacity, good castability, and low melting temperature and melting energy [1]. The die-casting process
stands out due to its excellent productivity, and the resulting potential to lower material cost through design strategies. Die-casting magnesium alloys can produce
acceptable strength values but the room temperature ductility remains fairly low. In addition, the problem of gas
porosity due to splashing during mold lling remains a
complex issue, which renders heat treatment to improve
the properties dicult [2].
For high-performance structural applications of magnesium alloys it is desirable to develop wrought magnesium
products, such as extruded proles, rolled sheets and forgings, which in contrast to die-cast products possess higher
strength and ductility. A number of technical issues, however, limit the wider application of existing wrought magnesium alloys. The most signicant problems are the
limited formability and strong anisotropy in mechanical
behavior, particularly at temperatures below 200 C, and
the low corrosion resistance. The corrosion performance
of Mg alloys depends strongly on the microstructure and
alloying elements, particularly on the impurity level of
heavy metals, such as Cu, Ni, Co and Fe. Optimizing the
corrosion resistance of Mg alloys requires knowledge of
the phases present in the alloy, their fraction and distribution, and most importantly their electrochemical compatibility with the Mg matrix. In MgAlZn (AZ) alloys, it
has been established that the b-phase (Mg17Al12) is cathodic to the matrix, and can either enhance or degrade the
corrosion resistance of the alloy depending on its distribution [3]. If it is uniformly distributed at the grain boundaries, forming a dense network, it functions as a
corrosion barrier and improves the corrosion resistance
of the alloy signicantly, but also decreases the ductility.
On the other hand, if the amount of b-phase in the matrix

820

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843

is not sucient, individual b-phase precipitates can cause


localized galvanic corrosion and increase the corrosion
rate. For MgAlMn (AM) alloys, the phase Al8Mn5 is
used to reduce the Fe and Ni content during casting, which
gives this alloy a reasonable corrosion performance. In Alfree Mg alloys, the corrosion behavior can be improved by
additions of rare earth elements and yttrium.
2.2. Ambient temperature formability characteristics of
conventional magnesium alloys
The limited formability of magnesium alloys at ambient
conditions represents a major challenge that prevents the
more widespread use of these materials. This limitation
leads directly to high production costs and low production
rates. For instance, conventional magnesium alloys are
generally extruded more slowly than aluminum alloys
because the temperature window where the material is
workable, yet does not suer incipient melting, is quite narrow. For the same reason, magnesium alloy sheet production has so far been restricted to elevated temperature
rolling. In contrast to aluminum and steel, which can be
nished cold, much of the reduction to nal gauges in magnesium has to be performed at elevated temperatures
between 300 and 450 C.
With respect to crystal structure, magnesium diers signicantly from its metallic lightweight competitor, aluminum, in having a hexagonal crystal structure. At ambient
temperature, plastic deformation of magnesium is limited
to
two
main
deformation
mechanisms,
basal
(0 0 0 1)h1 1 
2 0i slip and f1 0 
1 2gh1 0 
1 1i mechanical twinning. There are three dierent h1 1 
2 0i directions in the
(0 0 0 1) plane, and thus the magnesium crystal has only
three geometrical and two independent slip systems. On
the other hand, the aluminum crystal has 12 {1 1 1}h1 1 0i
geometrical and many combinations of ve independent
slip systems, and thus easily complies with the von Mises
criterion for homogeneous shape change [4]. Other slip systems, such as f1 0 
1 1gh1 1 
2 0i prismatic slip and
f1 0 
1 1gh1 1 
2 0i pyramidal slip, have the same hai-slip
direction in common but they require a larger critical
resolved shear stress (CRSS) for activation. Therefore, they
are usually harder to operate at room temperature, and at
the same time do not result in deformation out of the basal
plane.
Typical magnesium alloy sheet exhibits strong basaltype textures with grain orientations having basal planes
parallel to the sheet plane. Under loading in directions
either parallel or perpendicular to the sheet plane, basal,
prismatic and pyramidal hai-slip systems fail to accommodate any deformation because they all have a slip direction
parallel to the basal plane and the resulting shear strain in
all slip systems is zero. For such cases, slip vectors with a
component out of the basal plane are required. This can
be accommodated by slip on the f1 1 
2 2g-pyramidal plane
in the h1 1 
2 3i-direction, which is referred to as pyramidal
hc + ai-slip [5]. This slip mode oers ve independent slip

systems, thus satisfying the von Mises criterion for an arbitrary shape change. On the other hand it has a substantially
larger slip vector compared to hai-slip, and, hence, a markedly higher CRSS. Therefore, pyramidal hc + ai-slip in
magnesium usually requires substantial thermal activation,
i.e. high temperatures. With the absence of hc + ai-slip at
low deformation temperatures, the hexagonal crystals have
no means to accommodate the imposed strain along the
sheet normal direction by crystallographic slip. This causes
high stresses, modest work hardening, and, most importantly, premature brittle fracture.
In addition to hc + ai-slip, mechanical twinning on the
pyramidal f1 0 1 2g and f1 0 1 1g planes can provide a
shear component parallel to the c-axis. The most common
twinning mode in magnesium is the f1 0 1 2g extension
twinning. This twinning mode accommodates extension
along the c-axis, and hence for rolling deformation and a
basal texture this type of twinning obviously cannot be utilized. By contrast, f1 0 1 1g contraction twinning produces
a favorable compression strain along the c-axis; however,
in practice, while compression twins exist in conventional
magnesium alloys, they are scarce and do not contribute
to plastic deformation in a quantitative manner.
Like cold processing of rolled sheet, cold extrusion of
magnesium alloys faces similar diculties. During extrusion processing, the basal planes in the deformed material
are usually aligned parallel to the extrusion direction
(ED), forming a strong and homogeneous h1 0 
1 0i || ED
ber texture. Subsequent mechanical loading in compression along the ED will impose a tensile strain component
parallel to the c-axis of crystals and will, thus, massively
activate f1 0 1 2g extension twinning throughout the
strongly textured microstructure. As f1 0 1 2gh1 0 
1 1i
causes an orientation change of 86.3 h1 1 2 0i, the basal
planes of twinned grains will rotate by 86.3 from their original orientation, so that their c-axes become closely
aligned with the compression direction. This microstructural evolution normally occurs during the rst 6% of plastic deformation and causes a huge and unfavorable texture
change. Subsequent deformation needs to be accommodated by unavailable c-axis deformation mechanisms,
which severely limits the ductility of magnesium extrusions
for ED compression. For the reverse case of tensile loading
in the ED, the c-axis of crystals will experience compression, which triggers, however, only little f1 0 1 1g contraction twinning, and thus does not change the overall initial
texture, where the vast majority of grains are favorably oriented for prismatic hai-slip. However, the CRSS for prismatic hai-slip at ambient temperature is relatively high
compared to basal slip, which renders its activation somewhat dicult. This again contributes to the serious limitation of formability. In essence, the rapid texture formation
during deformation processing severely impacts the ductility and renders the material brittle.
One interesting approach to improve the cold formability of magnesium is to modify its hexagonal crystal
structure, either by reducing its c/a axial ratio or by

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843

completely transforming it into a cubic structure (e.g.


body-centered cubic (bcc)). This can be achieved by alloying with lithium, which also lowers the weight of the
already light magnesium even further (qLi = 0.58 g cm3).
With increasing lithium content the c/a ratio decreases
from 1.624 (for pure Mg) to 1.607 (for Mg17 at.% Li,
close to the solid-solubility limit) [6]. Between 17 and
30 at.% Li content MgLi alloys comprise a two-phase
structure consisting of the a-Mg rich (hcp) and the b-Li
rich (bcc) phases. The highly ductile b-single phase structure exists for Li contents greater than 30 at.% [7].
Although cubic MgLi alloys do not suer from ductility
problems, these alloys have high production costs, exhibit
low corrosion resistance and sometimes show low-temperature instability. Hexagonal MgLi alloys are also highly
ductile at ambient temperatures. This is attributed to a substantial increase in non-basal slip activity over entire
grains, particularly prismatic and pyramidal hc + ai-slip.
This is in contrast to pure magnesium, where the activity
of these slip systems is negligible at room temperature,
and at most limited to regions with large stress concentrations in the vicinity of grain boundaries. It is noted that the
enhanced non-basal slip activity is not only attributed to
the decrease in the c/a ratio resulting from lithium addition
but also to solid-solution softening eects that promote slip
on multiple planes [8]. By comparison to industrial Mg
alloys, binary MgLi alloys exhibit modest work hardening
behavior at room and elevated temperatures. They are also
very liable to grain growth at temperatures above 300 C,
which results in coarse microstructures and sharp recrystallization textures [9].
Since there is a strong correlation between the crystallographic texture evolution during thermomechanical processing, the operating slip and twinning mechanisms, and
the resulting mechanical behavior, it is obvious that controlling the microstructure (texture and grain size) has
broad potential to enhance the mechanical behavior of
wrought magnesium alloys. The principal goal of texture
modication for enhanced formability in magnesium is to
achieve a favorable alignment of the basal planes with
the deformation direction, whereas grain renement serves
to reduce the activity of twinning and promote additional
deformation mechanisms, such as grain boundary sliding.
A wide range of wrought magnesium alloy compositions
and processing schemes have been investigated, and it has
been reported that texture weakening/randomization was
possible in several cases with various alloy and processing
conditions. In the following we will review recent attempts
and discuss the underlying mechanisms, pinpointing
important questions that still need to be answered. We will
focus on conventional deformation processes, such as rolling and extrusion, and leave out non-conventional severe
plastic deformation methods, such as equal-channel angular pressing or high-pressure torsion, which despite their
great potential for grain renement, texture modication
and superplasticity [1012], have still to nd widespread
application and will be addressed in another contribution

821

of this issue. We will certainly not be able to give a comprehensive overview of all the recent developments in improving the formability of magnesium alloys, rather we will
focus on the critical issues, where substantial progress is
being made via intelligent texture modication. Finally
we will provide the reader with new perspectives that will
guide future magnesium alloy research and development.
2.3. Improvement of conventional magnesium alloys by
intelligent thermomechanical treatment
2.3.1. The role of random texture on ductility
The MgAlZn system is probably the most thoroughly
investigated of all magnesium alloy systems because it represents the most common commercial AZ alloys. With
respect to texture weakening and the improvement of formability of these alloys, several reports have been published
that focused on the key factors controlling the texture during deformation. For example, for the widely available
magnesium alloy AZ31, containing nominally 3 wt.% Al
and 1 wt.% Zn, a combination of extrusion and uniaxial
compression at 400 C and a constant strain rate of 10
4 1
s generated a random texture at strains >1 [13]. In
contrast, lower deformation temperatures and higher strain
rates resulted in sharp textures. Further investigations
showed that not only were the deformation parameters
important, but the choice of the starting orientation was
also crucial.
The material investigated was initially hot extruded and
depicted a typical h1 0 1 0i || ED ber orientation. Three
deformation orientations were explored, with the compression direction (CD) being set parallel (Fig. 1a), perpendicular (Fig. 1c) and aligned at 45 (Fig. 1e) to the ber axis of
the extrusion texture. This rendered the activation of slip
on basal and non-basal planes, as well as the activation
of extension twinning, easier for one starting orientation
than the other. When the material was compressed along
the ber axis of the extrusion texture, i.e. along the ED,
deformation was axisymmetrical for all grains and the
resulting nal texture at e = 1.4 was completely randomized, i.e. the texture intensity was less than 2 times random
(Fig. 1b). For the other orientations, though texture weakening was evident, the resulting textures were still well
dened and characterized by clustering of basal poles
around the CD (Euler angle, U, spread within 30, Bunge
notation) (Fig. 1d and f). Subsequent tensile testing of
the randomly oriented specimen at room temperature and
a constant strain rate of 104 s1 showed appreciable
enhancement of formability (ef  0.3) over the initial
extruded state (ef  0.18) with no decrease in strength
(Fig. 2). This shows that texture randomization is a promising avenue for the production of ductile magnesium
alloys.
Turning to the mechanism(s) responsible for randomizing the texture of AZ31 alloy, it is important to bear in mind
that the texture given in Fig. 1 is a result of high-temperature deformation comprising concomitant crystallographic

822

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843

Fig. 1. Schematic of the three tested orientations in uniaxial compression at 400 C and 104 s1 strain rate. CD, uniaxial compression direction; ED,
initial extrusion direction. (a), (c) and (e) represent the starting textures of each tested orientation with the CD parallel, perpendicular and 45 to the ED.
(b), (d) and (f) are resulting compression textures at e = 1.4. Texture representation in terms of orientation distribution function (ODF) with Bunge
convention using the MTEX tool box [14]. Texture intensity in this and all other gures is in multiples of a random distribution (MRD).

Fig. 2. Stressstrain curves of tension test at room temperature and


104 s1 strain rate comparing the maximum tensile elongation of an
extruded AZ31 specimen (Fig. 1a) and a uniaxially deformed specimen
with a randomized texture (Fig. 1b).

slip, dynamic recrystallization (DRX) and possibly grain


growth. Given the fact that the same deformation

parameters were applied in all three cases, there ought to


be discernible dierences in the deformation and recrystallization behavior that make the achievement of a randomly
textured material quite sensitive to the choice of starting
orientation. If we rst consider plastic deformation, it is
obvious that the activation scenarios of deformation mechanisms were dierent for each starting orientation. For
example, for the orientation where the texture was completely randomized, i.e. the applied force was parallel to
the h1 0 1 0i direction, prismatic slip, f1 0 1 2g extension
twinning and pyramidal hc + ai-slip were all potential
deformation mechanisms (thermally activated and geometrically favored). However, prismatic slip and mechanical
twinning can be readily excluded from playing an important
role since prevalent prismatic slip activity would only retain
the starting orientation, and twinning would cause a sudden
and large orientation change from Euler angle / values
close to 90 to values close to 0, which was not observed
in the recorded texture development as a function of strain,
reported elsewhere [13,15]. In this case, only pyramidal
hc + ai-slip is expected to dominate the deformation.
Despite its great advantage of providing ve independent

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843

slip systems, the activity of this slip nevertheless has to produce some texture component. The absence of pronounced
texture components, as shown in Fig. 1b, cannot be
achieved by any slip condition.
If we add DRX to the scenario, some degree of texture
weakening may certainly be expected. As seen, DRX was
evident in all three deformed orientations. The resulting
optical microstructures at e = 1.4 were fully recrystallized, and the average DRX grain size was comparable
for all three orientations [13]. It is acknowledged that some
recrystallization mechanisms, such as particle-stimulated
recrystallization, can generate a large spread of crystallographic orientations, and thus randomize the nal bulk texture by recrystallization and grain growth. However, for
the investigated alloy AZ31, only a few second-phase particles were present, so that any possible role of particles was
considered insignicant. Microstructure observations at
small deformation strains indicated that incipient nucleation of DRX occurred mainly through conventional serrations and bulging mechanisms of pre-existing high-angle
grain boundaries [1618]. In the course of DRX, with
increasing strain the deformed original grains were progressively consumed by the expansion of dynamically
recrystallized regions forming multiple layers of what is
known as necklace structure [19]. The rst necklace grains
formed by the bulging mechanism usually have orientations close to the orientation of the adjacent parent grains
[20]. Misorientations of 5 have been predicted to be common, and in some cases misorientations as high as 20 were
reported [21]. The expansion of the necklace structure was
shown not to proceed by means of repeated bulging on the
recrystallization front, but rather by other mechanisms that
lead to the loss of orientation coherency of the recrystallized volume with the deformed matrix [20]. In the case
of Mg, this phenomenon can be pictured as a continued
change in orientation of the DRX grains as deformation
progresses, up to large misorientations and a tendency to
texture randomization. Such a mechanism was proposed
in 1982 by Ion et al. [22] for Mg0.8% Al alloy. The
authors referred to the mechanism as rotation recrystallization owing to lattice rotations in the grain boundary
regions, which were also nucleation sites of recrystallization. They argued that lattice rotation during deformation
arises from the formation of subgrains along prior grain
boundaries, and there is a continuous increase of misorientation due to absorption of dislocations until a high-angle
boundary is obtained.
The question remaining here, however, is why the mechanism of rotation recrystallization contributed to full texture randomization in one orientation and failed to do so
in the other orientations. In a recent in situ study on
high-purity Al bicrystals subjected to external mechanical
stress at elevated temperatures [23], stress-induced grain
rotation concurrent with boundary migration was recorded
for a grain boundary with a mixed tilttwist character. In
polycrystals, most grain boundaries are of mixed type,
and depending on the stress direction with respect to the

823

grain boundary plane, the misorientation angle resulting


from stress-induced grain rotation may increase or
decrease. Therefore, there is good reason to believe that
the contribution of DRX-grain rotation to texture randomization is orientation dependent. For a favorable orientation (as in Fig. 1a) the fraction of suitably oriented grain
boundaries for stress-induced grain rotation during DRX
is most likely large enough to result in a complete loss of
texture. Additional investigations are necessary to provide
further support for this hypothesis.
In another texture study by Backx [24] on extruded
AZ31 combining heat treatment with uniaxial compression
at 250 C and 102 s1 up to dierent strains, texture randomization was recorded for 70% deformation strain
recrystallization annealing at 350 C for 30 min. This specimen demonstrated a remarkable ductility increase up to
e = 0.4 during subsequent uniaxial compression at room
temperature. The author attributed the enhanced mechanical properties to an optimized texture and microstructure
obtained by intelligent thermomechanical treatments.
Although the underlying mechanisms were not elucidated
in detail, it can be concluded from that work that static
recrystallization (SRX) can also randomize the texture in
AZ31. In essence, the degree of randomization will depend
on parameters such as the deformation strain, the grain size
and the texture of the annealing microstructure; hence the
attainment of optimal mechanical properties requires a
deep understanding of microstructure evolution during
processing.
2.3.2. The role of particle-stimulated recrystallization on
texture formation
Other investigations considered the eect of secondphase particles on the texture development in the
MgAlZn alloy system. For dierent aluminum contents,
texture formation during plane-strain compression (PSC)
at various values of Z was investigated [25], where Z is
the temperature-corrected strain rate, also
 termed the
Q
ZenerHollomon parameter; Z e_ exp RT
, where Q is
the activation energy and R is the gas constant. In addition
to varying Z, two initial rolling orientations were investigated, with the channel-die compression direction, CD,
being applied parallel to the previous rolling direction,
RD (frequently referred to as in-plane compression). The
extension direction was set in one sample parallel to the
previous sheet normal direction, ND, and in another sample parallel to the previous transverse direction, TD. This
rendered the activation of f1 0 1 2g extension twinning
easy for the former case and dicult for the latter. Owing
to the low Al content in the AZ31 alloy, the amount of bphase was negligible. In the AZ61 and AZ91 alloys, with
nominal Al contents of 6 and 9 wt.%, respectively, Mg17Al12 precipitates were primarily distributed at the grain
boundaries, and within the grain interior had average particle sizes of 510 lm (for AZ61) and 2030 lm (for AZ91).
During PSC testing at 200 C the volume fractions of the
second phase in the AZ61 and AZ91 alloys amounted to

824

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843

2.9% and 5.1%, respectively. At 400 C, the b-phase was


partially dissolved and the respective volume fractions were
1.1% and 1.7% [25].
The results showed that the impact of the b-phase precipitates on texture weakening during PSC at 200 C was
most pronounced for the AZ91 alloy relative to AZ61
and AZ31 (Fig. 3). This suggests that the precipitate density, and most likely also the precipitate size, are important
factors in magnifying the impact of particles on texture
development. The eect of Mg17Al12 precipitates on weakening the texture was qualitatively explained by enhanced
nucleation of DRX at second-phase particles. A suciently
large volume fraction of the b-phase causes nucleation of
grains with diverse orientations (Fig. 4). This view was corroborated by deformation experiments at 400 C and 10
4 1
s up to the same strain of e = 1, which yielded much
sharper textures (Fig. 3). At that temperature the b-phase

was mostly dissolved in the magnesium matrix and had


thus played little part during DRX.
2.4. Improvement of magnesium and its alloys by
microalloying with rare earth (RE) elements
2.4.1. Formation of soft textures in Mg-RE sheet and
extrusions
Magnesium alloys with rare earth (RE) elements have
been thoroughly investigated in Russia [26], and many indications were long ago found for the improved properties of
this class of alloys including grain renement, better formability at low temperatures, and enhanced strength and
creep resistance at elevated temperatures. In recent years,
wrought Mg-RE alloys have again attracted much scientic
attention since their deformation and recrystallization textures were found to be much weaker and less common than

Fig. 3. Schematic of the investigated sample orientation for the PSC tests and the corresponding starting textures of the three AZ alloys (middle row) in
terms of (0 0 0 2) basal pole gures. Top row corresponds to PSC textures at 200 C, 104 s1 strain rate and a true strain of 1. Outlined pole gure
indicates texture randomization. Bottom row corresponds to PSC textures at 400 C and same strain rate and nal strain as in the top row.

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843

825

Fig. 4. EBSD data of AZ91 specimen that showed a randomized PSC texture in Fig. 3. (a) Kikuchi band contrast (BC) map with high-angle boundaries
(>15) in black and low-angle boundaries (515) in yellow. (b) Inverse pole gure (IPF) map of the circled area in (a) and corresponding texture in terms
of discrete and contour data.

typical textures observed in conventional Mg alloy sheet or


extrusions.
In an attempt to eliminate the yield asymmetry in magnesium extrusions, Ball and Prangnell [27] were the rst to
note that a commercial WE43 alloy containing additions of
Y and RE elements can develop more random-type textures during extrusion, as opposed to conventional Mg
alloys. At that time, the randomized texture in WE43 was
attributed to particle-stimulated nucleation (PSN) of
recrystallization.
The fact that RE elements are relatively expensive and
not readily available throughout the world has promoted
considerable eorts to examine the potential of RE additions at microalloying levels and to check whether texture
modication is also possible when RE elements are in solution. In their study on sheet textures in MgZn alloys containing dilute additions (<1 wt.%) of RE mischmetal (a less
expensive mixture of RE elements) and Y, Bohlen et al. [28]
reported that the overall texture strength and the basal pole
intensity aligned with the sheet normal direction was lower
for RE-containing alloys than for conventional alloys. Not
only was the texture weaker, but the typical character of
the sheet texture was also altered, which had positive implications for sheet formability. Later, it was reported several
times in the literature that sheet textures of many RE elements containing Mg alloys cause a greater tilt of basal
poles in the sheet transverse direction, TD, rather than in
the rolling direction, RD (Fig. 5), which promotes more
activation of basal slip during loading in TD than in RD.
This causes a mechanical and plastic anisotropy response
in terms of in-plane yield strength, elongation to failure
and strain anisotropy (r-value) dierent from those
observed in conventional Mg alloy sheet material. The
TD-spread in RE sheet textures was found to reduce the
planar anisotropy in comparison with conventional alloys
that exhibit r-values between 2 and 4. Reduced planar
anisotropy, Dr, of rolled sheet was reported to enhance
the forming behavior under straining conditions, which is
required for sheet thinning [29].
The benecial eect of RE elements in magnesium alloys
has stimulated research on a wide range of alloys and

compositions involving dierent mixtures of RE elements


and Y. In a study to examine the eect of microalloying
addition of Y- and Nd-based mischmetal (Table 1) on
the extrusion texture and microstructure of MgZnZr
(ZK) alloys, appreciable texture and microstructure modication was achieved (Fig. 6). Fig. 6a and b show representative electron backscatter diraction (EBSD) micrographs
(inverse pole gure maps) of the extrusion microstructures
of both ZK10 (standard alloy version) and ZWEK1000
(microalloyed version with RE/Y) alloys, respectively,
and the corresponding misorientation angle distributions
(MADs) of the grain boundaries. Clearly, the addition of
mischmetal and Y aects the recrystallization behavior of
the material and weakens the texture. The large elongated
grains in the extruded ZK10 alloy were replaced by coarse
equiaxed grains in the RE/Y-containing alloy, which was
reected in the MAD by a pronounced reduction of the
misorientation peak at 5 and concomitant increase of
the misorientation peak at 30. The peak intensity at low
misorientation angles of 5 is an indication of the extent
of subgrain structures within the unrecrystallized elongated
grains. The misorientation peak at high-angle boundaries
of 30 is most probably due to a common 30[0 0 0 1] misorientation relationship during growth of recrystallized
grains in hexagonal materials (e.g. [3034]) and its intensity
can be correlated with the recrystallized volume fraction.
As evident from the orientation map in Fig. 6b, the
recrystallized grains in the modied alloy comprised a
broad spectrum of orientations that were dierent from
the sharp h1 0 1 0i || ED deformation orientation that dominates the traditional extrusion texture in the ZK10 alloy.
The wide variety of orientations of recrystallized grains
and their prevalent volume fraction in the microstructure
of the ZWEK1000 alloy was found to randomize the bulk
texture to a large extent. While this is still a subject of
current research it is proposed that stable second-phase
particles with a complex mixture of Zn, Nd, Y and Mg
are eective nucleation sites for DRX during extrusion at
400 C and give rise to randomly oriented nuclei. Fig. 7
shows good microstructural evidence of very ne
grains (d  5 lm) spotted in the immediate vicinity of

826

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843

Fig. 5. (a) Recalculated (0 0 0 2) pole gure and (b) basal pole density plot as function of tilt angle about the sheet normal direction, ND, showing an
example of RE sheet texture in a Mg0.9% Zn0.7% La0.2% Zr (wt.) alloy with obvious trend for TD spread opposed to the other sheet directions.

Table 1
Chemical composition of the investigated benchmark alloy ZK10 and the modied version containing Y and Nd-rich mischmetal.
Alloys

Zn (%)

Zr (%)

Nd (%)

Ce (ppm)

Gd (ppm)

Y (%)

Mg

ZK10 (standard)
ZWEK1000 (RE/Y modied)

1.51
1.47

0.41
0.38

0.20

31

11

0.19

Rest
Rest

Fig. 6. EBSD microstructures (IPF maps) along the extrusion direction, ED, and corresponding misorientation angle distributions for (a) conventional
alloy ZK10 (benchmark), and (b) modied alloy ZWEK1000.

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843

827

Mg7Gd6Y0.5Zr (wt.%) alloy [37], an unusual prismatic


texture component was observed, in which the c-axis was
aligned parallel to the extrusion direction, perpendicular
to the typical basal extrusion texture of magnesium alloys.

Fig. 7. Optical micrograph (ED plane) of the extruded ZWEK1000 alloy


showing one example of numerous ne-grained areas associated with
second-phase particles.

second-phase particles, whose size varied between 1 and


10 lm. Other grains, outside of the highlighted area in
Fig. 7 (where no particles were present), were 10 times
larger. While the role of particles in generating new recrystallized orientations may appear to account for a small volume fraction, recrystallizing grains seemed to readily
undergo grain growth, consuming a larger volume of the
extrusion microstructure, and thereby contributing to texture weakening.
From the above ndings it can be concluded that small
additions of Nd-rich mischmetal and Y to the extrusion
alloy ZK10 can signicantly weaken the extrusion texture,
yet the position of the texture peak was still close to the
h1 0 1 0i ber. This is not necessarily the case for other
extrusion alloys with additions of RE and Y. Stanford
and Barnett [35] reported a shift in the orientation peak
of extruded Mg alloy ME10 (1 wt.% Mn and 0.4 wt.%
Ce-rich mischmetal) from h1 0 
1 0i (seen in AZ31) or
h1 0 
1 0ih1 1 
2 0i (seen in M1) to a new position h1 1 2 1i
(about 30 from h1 1 
2 0i of the inverse pole gure), which
was termed in later studies the RE texture component. A
study by Huppmann et al. [36] on a similar alloy ME21
(2.1 wt.% Mn and 0.7 Ce-rich mischmetal) examined the
role of extrusion parameters (billet temperature, extrusion
ratio and cooling conditions) on the formation of the RE
texture component. In addition to the aforementioned
h1 1 2 1i component, they observed two additional RE texture components, h1 1 
2 2i and h2 0 
2 1i, parallel to the
extrusion direction. Moreover, it was shown that the variation of extrusion parameters and the cooling conditions
allows a variation in the extrusion texture. Both studies
concluded that the RE texture component is well oriented
for basal slip when tested in the appropriate orientation,
which results in a substantial gain of ductility and a reduction of the tensioncompression asymmetry typical for
conventional wrought Mg. While the focus of this paper
is on dilute MgRE alloys, it is noted that unusual textures
develop also in heavily alloyed MgRE extrusions with
more than 10 wt.% total RE. For example, in an extruded

2.4.2. Comparison of dierent RE additions with respect to


texture modication
In order to gain additional insights into the origin of RE
texture modication, recent eorts have focused on the
eect of single RE additions on the formation of deformation and recrystallization textures. Basically, dierent
RE elements can form dierent intermetallic compounds
[26,38] that contribute to distinct properties. Additionally,
in the form of solutes, individual RE elements may interact
dierently with dislocations and grain boundaries. In single
RE alloying studies, individual RE elements that are readily soluble (e.g. Gd, Nd and Y) or insoluble (e.g. Ce and
La) in Mg are usually chosen to study the role of RE in
the form of both solute solutions and two-phase alloys.
Investigations were mainly concerned with (a) examining
the potential of individual elements as texture modiers,
and (b) determining the amount of RE element required
to modify the texture. Based on ndings obtained for both
aspects, good progress has been made so far toward highlighting the underlying mechanisms for RE texture
modication.
2.4.3. The role of REsolute segregation
In a comparative study on microalloying Mg with dierent RE elements and Y for texture modication [39], it was
demonstrated that the minimum level required is dierent
for each RE element. At suciently high concentrations
between 300 and 600 ppm, MgLa, MgCe and MgGd
produced the RE texture component (h1 1 2 1i parallel to
ED). Strangely, Y did not show this texture type at any
of the concentrations examined (0.020.17 at.% or 0.08
0.62 wt.%). A similar trend was reported for extruded
MgMn alloys with single additions of Ce, Y or Nd [40],
where Y was a much weaker texture modier than Ce
and Nd.
EBSD investigations have shown the formation of the
RE extrusion texture to be associated with a high propensity for DRX nucleation at shear bands [41]. Furthermore,
it was proposed that the eect of individual RE elements on
the location and intensity of the RE texture component is
due to solute interaction with dislocations and grain
boundaries. Ce and La have larger atomic radii than Gd
and Y, and were found to show the strongest eect on texture and grain size [39]. In support of their hypothesis that
RE solutes have a strong interaction with dislocations and
grain boundaries, the authors conducted another study on
extruded Mg1.5 wt.% Gd alloy that went through a texture transition (h1 0 1 0i || ED ! h1 1 2 1i || ED) with
increasing extrusion temperature [42]. The aim of the study
was to investigate the correlation between solute segregation to grain boundaries and the RE texture eect. By
means of energy-dispersive X-ray spectroscopy (EDS)

828

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843

performed in a transmission electron microscope (TEM),


the authors reported a signicant increase in the local concentration of Gd at grain boundaries compared to the bulk.
Small amounts of other non-RE elements (Al, Mn and Cu)
were also present in the material but showed less segregation than Gd. The level of solute segregation was found
to depend on the extrusion temperature and correlated well
with the texture transition that occurred at 490 C. Temperature plays a role in the competing eects of solute
diusivity and grain boundary mobility. Elevated temperatures promote solute diusivity but decrease segregation,
and at suciently high temperatures the boundary migration rate becomes large enough for the grain boundary to
break free from its solute cloud and move freely [43].
Fig. 8 is a cast microstructure of a ZNK100 alloy (where
the letter N represents Nd) with 0.1 at.% Nd, showing good
evidence of preferential segregation of RE solute atoms to
grain boundaries. Apparently, the local concentration of
Nd at the grain boundary was in fact much higher than
its bulk concentration, leading to the precipitation of very
ne second-phase particles (<1 lm) at the grain boundary
upon homogenization treatment at 450 C for 12 h and
rapid quenching.
With respect to recrystallization, solute segregation may
have little inuence on the nucleation process but is known
to strongly retard the boundary mobility of recrystallization nuclei. This retardation of recrystallization is generally
observed in most MgRE alloys (e.g. [40,44]), and serves to
modify the recrystallization texture in various respects. On
the one hand, it can lead to increased formation of deformation heterogeneities with large crystallographic misorientations, such as shear bands and/or specic twin types
that lead to alteration of the dominant recrystallization
mechanisms (oriented nucleation aspect). On the other
hand, the segregation phenomenon can aect the growth
advantage of specic orientation relationships known for
their high boundary mobility, which can result in new
recrystallized orientations (oriented growth aspect). From
the current perspective, it is likely that both eects are

Fig. 8. SEM image of a cast Mg0.6% Zn0.6% Nd0.3% Zr (wt.) alloy


after homogenization treatment at 400 C for 12 h and water quenching
revealing extremely ne precipitates at the grain boundaries.

coupled and contribute strongly to the formation of unusual RE recrystallization textures. It should also be noted
that RE solute segregation can be coupled with RE solute
clustering in the matrix according to recent atom probe
tomography measurements [42]. Solute clusters could
prompt plastic strain heterogeneities similar to those
observed in systems with shearable particles, and could
thus certainly aect the nucleation and growth process of
recrystallization.
2.4.4. Impact of dierent RE elements on the mechanical
properties
While the addition of RE elements to thermomechanically processed magnesium primarily aims at weakening
the texture, the desired engineering outcome is improved
deformation behavior and enhanced property combination
of strength and room-temperature ductility. The link
between texture and formability has been well established,
and the connection between RE additions and texture is
currently being increasingly enlightened by intensive
research on the subject. Since ductility improvement in
magnesium could also benet from additional factors other
than the texture, such as grain renement, increased workhardening rate and strain-rate sensitivity, it is important to
determine the impact of RE elements on these parameters.
In a recent attempt to examine the eect of known texture modiers Gd, Nd, La and Ce-rich mischmetal on the
formation of sheet textures and the resulting tensile properties in a ZK10 alloy [45], it was found that the room-temperature tensile elongation at fracture was remarkably
improved over conventional alloy sheet, such as AZ31
(Fig. 9). From the investigated RE elements, Gd demonstrated the highest potential to randomize the sheet texture
(two times random, upon 75% hot rolling and recrystallization annealing at 400 C for 1 h). The other investigated
RE elements yielded somewhat sharper textures (four to
ve times random), characterized by a broad scatter of

Fig. 9. Modied sheet textures of ZK10 magnesium alloys after microalloying with dierent RE elements and the resulting tensile elongations to
fracture at room temperature and 5  104 s1.

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843

basal poles toward the sheet transverse direction. Good


correlation was found between the texture intensity and
maximum elongation to fracture, and also between the
hardening exponent n and the texture. The Gd alloy with
the weakest texture showed a remarkable increase of
room-temperature ductility up to 30% in all three sheet
directions. This was coupled with an average r-value of
1.025 and a planar anisotropy Dr of 0.05, which translates
into enhanced sheet thinning, less earing propensity and a
uniform strain distribution in the plane of the sheet. In
addition, the hardening exponent was increased up to an
impressive value of 0.45, which was found to eectively
inhibit the onset of localized deformation described by
the Conside`re criterion [45]. Regarding the recrystallized
grain size, there was no striking dierence in the dierent
alloys. The values were in the range of 2842 lm. The ultimate tensile strength was comparable to conventional sheet
but the yield point could denitely benet from some
improvement, obtained potentially by optimizing the alloy
composition and rening the grain size. It is challenging,
however, to increase the yield strength without sacrifying
ductility.
2.4.5. Eect of RE elements on the relative activities of slip
at elevated temperatures
Although there is general agreement in the literature
that texture modication by addition of RE elements
occurs primarily during recrystallization and grain growth,
the role of RE elements (in the form of both solute and particles) on the deformation mechanisms and the microstructure which precedes recrystallization remains important.
From a recent study [44] aiming at quantifying the deformation mechanisms during high-temperature plane-strain
compression of magnesium alloy ME20 containing cerium
as the main RE element compared to a benchmark alloy
AZ31, it was seen that despite the same deformation conditions (Z parameter) and qualitatively comparable initial
textures, the two alloys showed considerably dierent texture development (Fig. 10). This behavior was the result
of signicant dierences in the activation of deformation
modes, which was analyzed on the basis of experimental
texture development and texture simulations using an
advanced cluster-type Taylor model that accounts for grain
interaction [46].
As evident from Fig. 10, the texture development in the
conventional alloy AZ31 was very similar at both Z conditions, 200 C/102 s1 (high Z) and 400 C/104 (low Z).
From previous studies, it is established that the resulting
AZ31 textures at e = 1 are the outcome of prevailing prismatic slip activity with a small contribution from other haislip modes operating in some disoriented grains. On the
other hand, the RE alloy ME20 showed an interesting
trend of rotating the basal poles from their initial TD orientation toward the compression direction, CD, which
indicates a reduced activity of prismatic slip relative to
other slip modes, particularly during high Z deformation
at 200 C. This view was substantiated by simulation

829

results shown in Fig. 10 revealing that at 200 C pyramidal


hai-slip was much more important than prismatic slip; in
fact, at incipient deformation it was even more important
than basal slip that was impeded by the initial TD orientation. By increasing the temperature to 400 C, there was a
strong competition between prismatic and pyramidal haislip to accommodate the deformation. This competition
was extended to include also basal slip at strains larger
than 20%. Under such a multiple slip condition, pyramidal
hai-slip played an important compensating role preventing
the PSC deformation texture from turning into either a
basal or prismatic texture, which has positive implications
for ductility.
From the above ndings it seems that Ce addition
changes the relative CRSS of the various slip systems compared to benchmark AZ31, which points in the direction of
solid-solution hardening and softening eects on the dierent slip/twin mechanisms, even in the presence of particles
(micrometer-sized Mg12Ce and nanometer-sized pure Mn
particles). There are several reports in the literature on
the eect of non-RE elements such as Al, Zn and Li on
softening prismatic slip at the expense of basal slip [47
51]. However, a few studies on RE elements (e.g. [52])
reported that Gd alloying tends to increase the CRSS for
prismatic slip, which is consistent with the results presented
in Fig. 10 that show obvious strengthening of prismatic slip
in the high Z regime. On the other hand, when the temperature was increased to 400 C, solid-solution strengthening
of the prism plane seemed to diminish, probably due to
changes in the particle/solute concentration. The strengthening eect of prismatic slip did not seem to be coupled
with strengthening of pyramidal hai-slip.
It is noted that for the ME20 alloy, the addition of Ce
and the presence of nanometer-sized Mn particles strongly
retard DRX, and thus prevent it from playing a signicant
role in the texture evolution shown in Fig 10. The eect of
dierent initial grain sizes between AZ31 (40 lm) and
ME20 (15 lm) was not considered when discussing the
mechanisms responsible for the dierent texture development of the two alloys. However, based upon prior knowledge regarding the impact of grain size on texture
development, we do not suspect a dierence of 25 lm
to disprove the hypothesis of solute-related eects on the
deformation slip activity.
2.4.6. The eect of RE elements on the stacking fault energy
and hc+ai-slip activation at room temperature
As noted earlier, the addition of RE elements produces
weak recrystallization textures that promote basal slip,
which contributes to enhanced room-temperature ductility.
There are also reports in the literature that the addition of
RE elements in solid solution increases the ductility during
cold deformation without involvement of recrystallization
[5355]. A recent comprehensive study by Sandlobes
et al. [55] combined electron density functional theory
(DFT) calculations with TEM to quantify the fundamental mechanisms responsible for solute-related ductility

830

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843

Fig. 10. Comparison of the PSC texture development between a conventional AZ31 alloy and an RE-containing alloy ME20 during high (200 C/102 s1)
and low (400 C/104 s1) Z deformation up to e = 1. Texture prediction results (outlined) of ME20 oer an estimate of the relative CRSS values of active
deformation modes and their relative activities as a function of strain.

enhancement during cold deformation of MgY alloys.


The authors reported that both methods (TEM and
DFT) revealed a signicant decrease of the intrinsic stacking fault energy (SFE) with the addition of Y, and concluded that the reduced number of intrinsic stacking
faults enables the formation of dislocation structures on
pyramidal planes, and acts as a heterogeneous nucleation
source for hc + ai-dislocations, which correspondingly promotes the activation of hc + ai-pyramidal slip at room
temperature.
2.4.7. The role of non-RE elements on RE texture
modication
Alloying with RE elements is generally done using either
pure magnesium, resulting a simple binary MgRE alloys,
or dilute magnesium-based alloys, such as MgMn, Mg
Zn or MgZnZr, resulting in more complex alloy systems.
While this has recently been intensively studied, little attention has been paid to the role of accompanying non-RE
elements on the RE texture modication. Preliminary
results indicate that the texture weakening eect of RE elements is substantially magnied by the addition of other
non-RE elements, such as Zn and Zr.

Figs. 11 and 12 show interesting comparisons between


binary MgRE alloys and their corresponding quaternary
systems with added Zn and Zr. Fig. 11 presents the cast
microstructures of Mg1.07% Gd (wt.) (G1) and Mg
0.92% Zn1.04% Gd0.57% Zr (ZGK110) alloys, and the
resulting textures after rolling at 400 C and thickness
reduction of 80%, followed by annealing at 400 C for
1 h. It is evident that the strongest texture weakening
occurred in the ZGK110 alloy during recrystallization
annealing (Fig. 11b). The corresponding texture of the binary G1 alloy revealed a considerable reduction in basal
pole intensity of one texture component (compared with
the rolling texture) but the total maximum intensity was
still high. The combination of non-RE elements (Zn and
Zr) with Gd seemed to be important also for the development of an RE-sheet texture with TD-spread during rolling, whereas the addition of only Gd to magnesium
resulted in a conventional double-peak basal texture with
high intensity. Although these are preliminary results, it
is obvious from Fig. 11 that the formation of RE textures
and the associated texture randomization during recrystallization depend strongly on grain size and the nature of the
second-phase particles, such as composition, size, shape,

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843

831

Fig. 11. Comparison of the cast microstructures and second-phase precipitation, and the resulting rolling and annealing textures between (a) binary Mg
1.07% Gd (wt.) (G1) and (b) Mg0.92% Zn1.04% Gd0.57% Zr (ZGK110) alloys.

Fig. 12. Comparison of the sheet texture development in rolled binary MgCe alloy (top) and a quaternary version of the same alloy containing
additionally Zn and Zr (bottom) as a function of annealing temperature.

distribution, etc. Another important aspect, not directly


seen in Fig. 11 is the nature of solutes and their interaction

with each other and with microstructural features, such as


dislocations and grain boundaries.

832

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843

Another example is given in Fig. 12 for annealing texture modication as a function of annealing temperature
in a binary Mg alloy containing 1.15% Ce and a similar
version containing in addition 1.05% Zn and 0.57% Zr
(wt.). While texture weakening was evident in both alloys
upon 1 h annealing at 350 and 400 C, it was much more
pronounced in the quaternary alloy containing Zn and
Zr, where the sharp rolling texture (10 times random)
was strongly randomized (2 times random). Notably,
for annealing temperatures up to 300 C, both alloys exhibited similar annealing textures, which were not much dierent from typical basal textures of conventional magnesium
sheet. With this information at hand, future investigations
will focus on annealing temperatures between 300 and
350 C to examine recrystallization and detect changes in
the microstructure and phase relations between the two
alloys in order to improve our understanding of the role
of non-RE elements in texture randomization during
recrystallization.
3. Aluminum alloys
3.1. Advantages and applications
In contrast to magnesium, the slightly heavier element
aluminum has already established a leading role in a wide
range of applications, due to its light weight, relatively easy
fabrication and attractive mechanical properties. The use
of these alloys started with their spectacular development
in the aerospace industry as soon as high-strength variants
were developed and became available in sucient quantities, more than 100 years ago [56]. In the past few decades,
increasing amounts of these alloys have also been used in
automotive applications due to the creation of robust
and easily applicable variants with good strength, formability, crash and corrosion performance [57].
3.2. Aluminum sheet alloys for automotive applications
The two main alloy systems used in automotive applications are the non-heat-treatable AlMg (AA5xxx) and the
age-hardened AlMgSi (AA6xxx) alloys, which show a
good combination of sucient strength and good formability [5762]. For specic applications (e.g. in bumpers and
crush-zone elements) the high-strength AlZnMgCu
(AA7xxx) aerospace alloys are also being used mostly
as extruded parts. However, due to limitations in corrosion, joining and age-hardening characteristics, this and
the other high-strength AlCuZnMg (AA2xxx) alloy
group are less suitable for conventional mass-produced
automotive parts. However, as well as alloy additions, thermomechanical processing is of major importance [63]. This
implies close control of texture, which can be quite strong
in Al alloys [64] and thus signicantly inuences anisotropy
and related key properties, such as yield locus and forming
behavior [65,66].

The strength and formability of the main group of nonheat-treatable AlMg (5xxx) sheet alloys for automotive
applications is based on the mechanism of solid-solution
hardening by Mg additions, usually up to 5% [58]. This
also causes a high strain-hardening eect that helps to stabilize the sheet during stretch forming, avoiding local necking. Thus high-Mg-containing AlMg alloys enhance both
key properties, i.e. strength and formability.
3.3. 5xxx AlMgMn alloys
The use of 5xxx AlMg alloys is well established in chassis and various structural applications, due to their good
forming behavior. They are also used in sheet panels, but
seldom as exterior parts due to sheet surface irritation of
Luders lines, caused by the PortevinLe Chatelier eect.
At room temperature and slow strain rates an avalanche
type dislocation motion occurs in AlMg alloys due to
the strong dislocation interaction (pinning) with diusing
Mg atoms.
AlMg alloy strength is mostly dened in terms of a
soft-annealed O temper. The signicant strength contribution due to strain hardening is seldom used in automotive parts (in contrast to beverage cans [67]) due to heat
treatments during processing (e.g. paint baking) and the
softening eects involved in any use phase at elevated temperatures. If exposed to elevated temperatures for long
times medium (max. 3%) Mg-containing alloys are used
due to the eect of intercrystalline corrosion (IC) in corrosive environments. In such cases for alloys containing
(>3%) Mg, special precautions in either processing or in
application conditions are required, while for predominant
moderate-temperature applications (e.g. in marine environments) high (5%) Mg alloys are common. In recent years
new variants of AlMg alloys with higher Mg contents (e.g.
3.5%) [69] or with small additions of Cu have been investigated showing improved properties, some of which are
used for body-in-white applications by Japanese automobile companies [68].
3.4. 6xxx AlMgSi alloys
6xxx AlMgSi are the well-established age-hardening
sheet alloys used in many automotive parts, including exterior panels, which have high requirements for surface
appearance. They also provide a high eective strength in
the age-hardened T6 condition, and lower strength and
good formability in the T4 (solution annealed and aged
at room temperature) condition [58,70]. Their characteristic strength evolution occurs following an additional heat
treatment which is applied in car production after forming,
in some cases simply during paint baking cycle of the bodyin-white. In the temperature range around 185 C a significant increase in strength occurs by precipitation hardening
when the sheet was processed by a high solution anneal
(>540 C) with fast quench in a continuous annealing line.

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843

The age-hardening behavior can be further enhanced by a


special pre-treatment pre-bake (avoiding the specic
retrogressive eect observed after long-term roomtemperature aging) and by deformation applied during
the production of parts, thus achieving sucient strength
after part forming and paint baking [70].
3.5. Other alloying additions
Mn has a benecial inuence on the mechanical properties and is added mainly to control grain size by forming
sub-micrometer-sized particles [71]. The microstructural
changes involved during annealing often include texture
eects since particles modify textures by changing the
recrystallization mechanisms (enhancing PSN [72]) and/or
inhibiting grain boundary mobility. This is especially
important in AA6xxx alloy sheet due to the required high
solution annealing temperatures that otherwise induce
grain growth. The corresponding recrystallization mechanisms can most eectively be controlled by choosing the
appropriate processing routes, especially the heat treatments, starting with ingot pre-annealing and homogenization [73,74].
Other elements, such as Fe, that must also be considered
in industrially processed alloys, are usually present as
impurities. Due to their low solubility they can have a negative eect on formability by forming large constituent
particles during eutectic solidication. They are the source
of local cracking during forming operations. Proper ingot
homogenization cycles, additions of Mn and large rolling
reductions aect their form and size and help to reduce
their negative inuence, as well as aecting texture formation during intermediate and nal annealing treatments
[73,74].
3.6. Forming characterization and simulation
In simple characterization of sheet material behavior
formability is often expressed by parameters such as
strain-hardening coecient n-value and anisotropy coefcient r-values, which are easily measurable in simple
tensile tests and were therefore used in early forming simulation software. Advanced forming simulation methods are
needed, however, for anisotropic materials (such as aluminum or advanced steel) which require more detailed (up to
16 or even more) parameters to fully describe the material
ow under various conditions [66]. These models were
developed for aluminum sheet where some anisotropy
eects need to be considered due to complex textures. Conventional interstitial-free steels, in contrast, show rather
simple ber textures with rotational symmetry in the sheet
normal [75], which is easily described by classical Bishop
Hill continuum models. For new steel grades (e.g. TWIP/
TRIP steels), however, the advanced models are now being
applied since directional eects play a major role in their
specic forming behavior. New crystal plasticity-based
codes have also been developed and applied in order to

833

include detailed texture data to describe Taylor-type


crystallographic slip and predict related anisotropy eects
[76].
3.7. Textures in aluminum alloy sheet
3.7.1. Rolling textures in aluminum sheet
During cold rolling of any aluminum alloy a typical fcc
rolling texture develops, which is composed of two bers
[77]: (i) the a-bre (= h1 0 0i parallel to the normal direction, ND, which connects the {0 1 1} h1 1 2i B orientation
with the {0 1 1} h0 1 1i G (Goss) orientation) occurring at
low strains; and (ii) the b-bre which aligns the main three
texture components C (copper), S and B (brass). Both
bers show some characteristic scatterings and peaks,
depending on the initial (casting or recrystallization) texture, and thus on the preceding processing history.
In nal customer operations the fully annealed tempers
are the most widely used conditions (except for beverage
can production [67]). Hence, the corresponding recrystallization textures are usually more relevant, and aect the
customer forming processes involved. The recrystallization
textures of rolled and annealed aluminum alloys have been
investigated in detail and fully described (e.g. [78,79,84]) by
advanced texture evaluation methods (X-ray diraction,
EBSD).
3.7.2. The Cube texture
The most prominent fcc recrystallization texture is the
classical Cube component (Fig. 13), which appears in most
hot-rolled Al-alloys with characteristic shifts or scatterings
of the orientation peak [78]. After cold rolling and annealing
it is dominant in high-purity Al alloys. However, also in nonheat-treatable Al-Mg (AA5xxx) and in age-hardenable
AlMgSi (AA6xxx) sheet alloys this component can nucleate
and grow signicantly during hot rolling and fast and hightemperature heating of the required nal solution annealing
treatment, aecting strength and formability [80].
The Cube texture originates from a classical nucleation
mechanism of metastable orientations existing as small
bands in the highly rolled microstructure. In hot rolling
they gain in stability due to the activation of non-octahedral slip systems. These Cube bands are surrounded by
divergent orientation zones which rapidly enhance the
local orientation gradient needed for sucient grain
boundary mobility. Growth is then strongly accelerated,
supported by a preferred oriented growth eect due to a
40 h1 1 1i orientation relationship to the surrounding
main rolling texture (b-bre) orientations, including all
of its variants: due to its high symmetry the Cube orientation can compensate minor deviations from this relationship and resistant grains can easily be by-passed
when surrounded. This explains why the Cube texture is
the best compromise orientation according to recrystallization simulations based on statistical orientation relationships, including preferred nucleation and oriented
growth eects.

834

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843

Fig. 13. Cube texture component in Al alloys. (a) Hot-rolled AA5182 AlMg (full ODF). (b) Cold-rolled and annealed AA6016 AlMgSi (u2 = 0
section only). (c) Cube and R texture in 95% rolled commercial-purity Al (AA1145). (d) ND rotated Cube and P texture in age-hardenable AlMgSi
(AA6111).

3.7.3. The R texture


Another prominent recrystallization texture component
in rolled and annealed Al sheet is the R component
(Fig. 13c), which is very similar to the S rolling texture
component, indicating either an in situ recrystallization
or a strain-induced boundary migration (SIBM) nucleation
process with subsequent growth [81]. It mainly occurs in
Fe-containing (AA1xxx and AA8xxx) alloys and is usually
not observed in rolled and recrystallized AlMg (5xxx) and
AlMgSi (6xxx) alloys used as automotive sheet. It thus

can be concluded that the recrystallization mechanisms


related to the R texture are not dominant in these alloys.
The R texture is interpreted as a retained rolling texture
S component, but often shows distinct dierences in intensity and exact position. If Cube nucleation is suppressed
early (e.g. by simultaneous precipitation eects) or runs
out of nucleation sites (e.g. at extreme rolling reductions
as in foil rolling) and no other nucleation sites are active
the R-orientation emerges and can grow substantially
[82]. It is assumed to nucleate from existing orientations,

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843

e.g. by a bulging SIBM mechanism at existing grain boundaries. It can grow preferentially by a 40h1 1 1i relationship
into one of its own symmetrically equivalent variants. It
then shows systematic deviations from the original S orientation, either due to preferred nucleation eects (near the C
orientation) or preferred growth eects (near the 40h1 1 1i
rotated S orientation) [81].
3.7.4. PSN recrystallization textures in aluminum
In industrial rolled and recrystallized AlMg (5xxx) and
AlMgSi (6xxx) alloys recrystallization nuclei often originate at deformation inhomogeneities, i.e. particles or shear
bands [80]. The latter is the Q component (near
{0 1 3}h2 3 1i) which was observed to originate at shear
bands as observed in highly cold-rolled AlMg alloys
[83]. The rst is the P (PSN [72]) component (near {0 1 1}
h1 2 2i) originating from local deformation inhomogeneities
around second-phase particles. Both mechanisms lead to
signicantly lower texture intensities, as compared to Cube
and R textures (Fig. 13c and d), but certain texture eects
can be observed and can help to reveal the underlying
mechanism.
PSN is an important recrystallization mechanism in
industrial aluminum sheet, inuencing textures mainly by
generating a nearly random orientation distribution. However, some characteristic texture eects can be observed in
the weak recrystallization texture (as shown, e.g., in
Fig. 13d) revealing a peak in {0 1 1} h1 2 2i P and in
{0 0 1} h3 1 0i ND rotated cube; the main peaks are very
close to orientation accumulations predicted from the rolling texture by a complete 40 h1 1 1i transformation. This
texture is interpreted as a preferred growth selection process taking place out of the large spectrum of potential
nuclei around particles, e.g. in early stages of recrystallization (microscale growth selection).
In heat-treatable AlMgSi AA6xxx alloys recrystallization is governed by PSN but this also competes with other
nucleation mechanisms, in particular with the Cube nuclei.
Hence the nal texture can vary and depends on the relative amounts of nucleation at either orientation. This is

835

illustrated in Fig. 14ac where the nal recrystallization


texture is inuenced by the homogenization during the initial ingot preheating treatment, with other process parameters kept constant [79,80]. In Fig. 14 the exact cube
orientation together with its RD scatter increase at the
expense of the PSN (e.g. P) orientations with increasing
homogenization treatment. The recrystallization textures
in AlMgSi alloys consist of the Cube orientation with
some RD and/or ND scatter and/or the P orientation
(and some R orientation) in dierent proportions depending on the particles formed during early processing. Here
also the supersaturation level plays a role which can inhibit
early growth of preferred (Cube) nuclei by simultaneous
ne precipitation. The decrease of cubeND and P (the R orientationnot shown in Fig. 14appeared to scale with
the Cube orientation) reveals a decrease in the eciency
of PSN with increasing precipitation of Mg2Si particles
during hot rolling and/or nal annealing. These nely dispersed particles tend to impede PSN more strongly than
nucleation at the Cube bands. Hence, nuclei emerging from
the Cube bands (i.e. Cube- and cubeRD-oriented grains)
and the grain boundaries (i.e. R-oriented grains) will prevail in the recrystallization texture [80].
3.8. Texture eects on mechanical properties
Fig. 15 illustrates the resulting plastic anisotropy in
terms of the variation in r(a) (Fig. 15a) and the (symmetrized) earing proles h(a) of deep-drawn cups (Fig. 15b)
for the three dierent materials. The texture variations have
a strong impact on the resulting plastic anisotropy since the
PSN textures (CubeND and P orientations, e.g. homogenization states C-1 and C-2, see Fig. 14a and b) generate maxima for both earing and r-values at an angle of about 45 to
the RD and minima at 0/90. Theoretical r-values were
predicted from the texture data [73] to give the complete
r(a)-curves tted to the experimental points in Fig. 15a.
Texture control can balance the pronounced anisotropic
properties of the prevalent Cube-recrystallization texture
with the opposite 0/90 ears and minimum r-values at

Fig. 14. Particle-stimulated nucleation (PSN) texture in cold-rolled and annealed AA 6010 AlMgSi alloy sheet with dierent ingot pre-annealing
processes.

836

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843

Fig. 15. Plastic anisotropy of the textures shown in Fig. 14: (a) r-values (symbols: experiments; lines: simulated from textures); (b) earing proles
(experimental).

45. The resulting texture variation plays an important role


in the material properties since their opposite eect on
anisotropy can be used to improve the formability of Al
MgSi autobody sheet.
3.8.1. Strength
Any variation in crystallographic texture has an immediate impact on material strength due to the changes in
geometry of the dierent slip systems activated upon ongoing or subsequent deformation. The simplest explanation
for this can be given by assuming single slip only on the slip
system with the highest orientation factor m (= 1/Schmid
factor l = cos k  cos v, with k and v being the angle
between the tensile and slip plane normal and the slip direction, respectively) in a tensile test where the variation can
be up to 84% (m varies from 3.67 for the h1 1 1i direction
to 2 for a central direction with k = v = 45), neglecting
any orientation-dependent strain-hardening eects during
deformation! For the other extreme of multiple slip (a minimum of ve slip systems) as dened by the Taylor model
the variation still can be as high as 61%. Any slip system
activation will occur within the limits of these two models,
and hence the variation range in strength can be estimated
accordingly.
3.8.2. Formability
The Cube texture is typical for all aluminum alloys
recrystallizing during hot rolling. In a fully recrystallized
AlMgMn hot strip the Cube texture causes a characteristic anisotropic behavior, e.g. as strong 0/90 earing in a
cup deep drawn from a circular blank as shown in Fig. 15b.
It is technically quite important, e.g. for AlMgMn can
body stock, where a strong hot strip Cube texture is
required for low earing and good formability of the coldrolled nish gauge sheet [67].
Full or signicant partial recrystallization in the preceding rolling stepseither during hot rolling or by intermediate annealingconsiderably reduces the nal Cube
texture strength. This means that a large recrystallized
fraction during hot-rolling passes reduces Cube texture

strength [71], while a short interpass times (as in multistand hot-rolling lines) or reduced temperatures increase
it. However, minimum Cube texture is needed when hot
strips are used for forming operations and a more isotropic deformation behavior is required, e.g. for the deep
drawing of wheel disk fabricated from an 8 mm AlMg3
hot strip. Here optimum formability is provided by controlled partial recrystallization where some of the hot-rolling texture is preserved to eectively balance the opposite
anisotropy of the Cube recrystallization texture. An
almost even cup rim prole can be produced which otherwise may lead to severe deviations in the rotational
symmetry.
To describe and compare sheet formability properties,
forming limit diagrams (FLDs) are used [65]. The maximum strain achieved under all biaxial strain states
depends on the strain hardening (e.g. n-value) and local
ow pattern, including the anisotropic material ow. It
is evaluated from tests under dierent loading/straining
conditions. In recent years methods have been developed
to generate FLDs from their principal determining factors
[85]. The simulation shows how strain hardening aects
the level of the FLD curves and texture modies their
shape.
Fig. 16 shows the results of an industrial AA6016 sheet,
processed in two dierent (proprietary) ways, revealing a
pronounced formation of a PSN texture with a signicant
increase in formability (6016-A30+) [86]. Another
important aspect of the formability of AlMgSi AA6016
sheet (also encountered in steel sheet forming) is the formation of uneven surfaces during sheet forming (called roping). As indicated in Fig. 17, a localized band-like
formation of many ne-grained Cube orientations is
responsible for this eect. Systematic process variations
are applied to control the texture in order to reduce the
volume and uneven distribution of Cube grain orientations and thus avoid roping. Both cases demonstrate the
importance of texture control in Al sheet processing to
ensure quality and enhance performance for automotive
applications.

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843

Fig. 16. Textures of AA6016 autobody sheet with processing variations.


(a) Conventionally processed AA6016. (b) Advanced processed AA6016.

Fig. 17. EBSD of AA6016 surface showing aligned recrystallized grains in


cube texture.

4. Summary and concluding remarks


The common features and signicant dierences of texture formation during sheet rolling and annealing have
been analyzed for the two most important lightweight metals used in automotive applications. Conventional thermomechanical processing of aluminum and magnesium alloys

837

produces sharp ber textures, which in fcc aluminum accumulates in the b-ber with characteristic peaks, depending
on the preceding processes and initial textures, whereas for
magnesium the basal plane rapidly aligns parallel to the
direction of primary material ow. While aluminum
deforms by classical slip on {1 1 1}h1 1 0i slip planes, easily
accommodating strain incompatibility in any orientation,
mechanical twinning in magnesium involves only the c-axis
deformation mode at room temperature with drastic consequences for texture and resulting mechanical performance.
For standard Mg sheet or extrusion textures, subsequent
tensile or compressive loading parallel or perpendicular
to the predominant ow direction requires deformation
along the c-axis, which under ambient conditions means
accommodation by mechanical twinning. For c-axis contraction, f1 0 1 1g twinning only activates at very high
stresses at which void nucleation occurs. For c-axis extension, f1 0 1 2g twinning is not very helpful, as it hardens
the texture by abruptly rotating it into contraction along
the c-axis. The diculty of forming magnesium at temperatures below 150 C is, therefore, attributed not only to a
lack of respective deformation modes but also to the consequence of activating twinning, particularly under c-axis
contraction.
As highlighted in this paper, attempts to enhance the
ambient formability of magnesium alloys have aimed at
lowering the stress for non-basal slip activity, by alloying
with Li for example, or by combining texture optimization
and grain renement obtained by conventional or
non-conventional processing. Additions of RE elements
have shown promise in this regard. Yttrium addition, for
example, has a ductilizing eect on ambient formability
that was attributed to enhanced pyramidal hc + ai-slip
activity associated with a signicant decrease of the intrinsic stacking fault energy. Another favorable eect of RE
elements is on texture modication that seems to hold for
mischmetal additions or single additions of Ce, La, Nd,
Gd, etc. The level of texture modication depends strongly
on the RE concentration, the processing parameters and
sometimes on other non-RE alloying elements, such as
Zn. Modied textures can show a decrease in texture intensity up to random level or exhibit RE-texture components that depart from standard sharp sheet and
extrusion textures. The weaker textures produced in these
cases are better aligned for basal slip, which results in
enhanced ductility and reduced mechanical anisotropy.
As a secondary eect, addition of RE elements also renes
the grain size of the alloy, which confers an additional ductility benet.
In rolled and annealed aluminum alloys the eect of
recrystallization generates a spectrum of new orientations
and texture components. Their interpretation and the
methods derived to control them for achieving advanced
properties in practical applications are highly developed
and industrially applied. The classical Cube component
occurs in many hot-rolled, as well as cold-rolled and
annealed Al alloys, in some cases with characteristic shifts

838

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843

and/or scatterings. It is a major inuence on many anisotropic properties, including forming behavior and related
surface eects. The R-component is very similar to the
main S rolling texture component indicating either an
in situ recrystallization (i.e. extended recovery) or SIBM
nucleation processes with subsequent growth. In industrial
alloys second-phase particles play a major role, modifying
the Cube texture and generating the P-component. With
variations in particle size and distribution and in the solute
content of ne precipitating elements suppressing the early
Cube nucleation dominance, the textures in AlMgSi
alloys can be systematically varied and properties modied
and formability signicantly improved. Other weak components originate at deformation inhomogeneities such as
shear bands, typical for weak recrystallization textures of
highly cold-rolled and annealed high-Mg-containing Al
Mg AA5xxx alloys.
For magnesium alloys containing RE elements, recrystallization plays a signicant role in modifying the deformation texture. With the employment of modern
characterization techniques, such as electron backscatter
diraction, atom probe tomography, high-resolution
energy dispersive X-ray spectroscopy and rst-principles
modeling, the mechanistic understanding of deformation
and dynamic recrystallization in correlation with RE elements has been markedly improved. Numerous mechanisms based on solid-solution and particle eects have
been suggested to be primarily responsible for the RE texture modication. The most important solute-related
mechanism involves solute segregation to grain boundaries
that serves to retard recrystallization, which has crucial
eects on the nucleation characteristics (e.g. the spectrum
of orientations that can nucleate or the type of nucleation
sites) and the relative mobility of dierent boundaries.
PSN is also observed in magnesium alloys, and seems to
be particularly promoted in the those containing RE additions. This mechanism provides more randomly oriented
nuclei that weaken the sharp rolling or extrusion textures.
It may appear to involve small volume fractions of the
microstructure, and be hence considered less important,
but in fact the orientations concerned may dominate the
bulk texture following recrystallization and grain growth.
In many cases, where both solutes and particles are present, the RE-induced texture change can be expected to be
impacted by several mechanisms at once. While the role
of recrystallization has been suciently highlighted by
many authors, the role of deformation in nucleating some
mechanisms for texture modication prior to recrystallization should also be considered important. In particular,
whether the deformed microstructure allows better development of deformation heterogeneities, such as shear
bands or other strain localizations that contain larger crystal rotations, or whether the twinning behavior shows any
distinctions from traditional Mg alloys regarding certain
twin modes (e.g. compression twins becoming equally
favorable with tension twins when loaded accordingly).

The favorable eect of some RE elements in providing multiple slip conditions can also have an impact on the nucleation of grains with distinct orientations.
It is anticipated that microalloying of RE elements combined with optimized processing for grain renement and
soft textures are going to help future conventionally processed wrought magnesium alloys become an important part
of automotive applications. It is, however, noteworthy that
in some cases, and after a careful choice of many parameters,
soft textures can be achieved with traditional (non-RE)
alloys, such as AZ31 or AM50. We showed that a combination of extrusion and uniaxial compression at 400 C in the
extrusion direction succeeds in randomizing the texture to
a large extent, which demonstrated positive implications
for subsequent ambient forming for this class of alloys.
In both magnesium and aluminum the formation of
recrystallization texture is controlled by the limited nucleation events and/or by a subsequent process of growth
selection out of the corresponding orientation spectrum.
In both materials specic alloying elements can be added
and thermomechanical processing modied so that they
aect the basic recrystallization mechanisms and allow
the texture, and thus the nal properties, to be varied.
The principal texture eects on product properties show
variations in:
 strength, either by the orientation dependence of slip or
indirect by changing slip modes, e.g. during age
hardening;
 formability, where deformation geometry is aected, e.g.
in the form of the r-value in tensile deformation or cup
height and ange thickness in deep-drawing operations.
Since property control is a key issue in many practical
applications, textures and related process variables are
accordingly being industrially controlled. Furthermore, if
analyzed carefully, useful information can also be obtained
about the underlying physical processes involved in industrially processed material. These two aspects combine both
the fundamental and the practical aspects of texture
research for material control and process optimization in
industrial applications.
Acknowledgments
T.A.S. would like to acknowledge nancial support
from the Deutsche Forschungsgemeinschaft (DFG), Grant
No. (AL 1343/1-1), and D.A. Molodov for helpful discussions. T.A.S. also thanks Indranil Basu, Feng Jiao, Xiaohui Li and Konstantin Molodov for their eorts in the
reported results. The assistance of Arndt Ziemons in producing magnesium alloys is sincerely appreciated. J.H. is
grateful to Olaf Engler and other former coworkers at
the Institute of Physical Metallurgy and Metal Physics
(IMM) at RWTH Aachen University and current
colleagues at Hydro Aluminium R&D Center, Bonn for

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843

many years of fruitful co-operation and many helpful


discussions.
Appendix A. Representation of texture in cubic and
hexagonal materials
Texture is a collective term for a non-uniform distribution of crystallographic orientations in a polycrystalline
aggregate [87]. A sample in which crystallographic orientations are fully random is called a random texture or is
often said to have no texture. If the crystallographic
orientations are not random, but have some preferred orientation, then the sample has a pronounced texture. In
texture measurement using X-ray or electron beam diffraction, the volume fraction of a particular family of
crystallographic planes with respect to a sample coordinate system is quantied. The axes of a sample coordinate
system are normally chosen according to important directions associated with the external shape change of the
specimen. For rolling deformation those are usually the
rolling direction (RD), the normal direction (ND) and
the transverse direction (TD). The texture obtained in this
way is an average value of the whole sample volume (provided that that the measured sample volume is statistically
large enough).
In order to understand how crystallographic textures are
formed and how they aect material properties, it is essential to have a method of representing and characterizing
them. Data obtained from texture measurements can be
presented graphically in dierent ways. The most important
ones are pole gures (PFs), inverse pole gures (IPFs) and
orientation distribution function (ODFs) [87,88].
1. Pole gure. PFs are two-dimensional stereographic plots
that describe the relation between the orientation of the
crystallographic axes of a crystal with respect to a coordinate system of a sample. An (h k l) pole gure shows
the distribution of the {h k l} poles in the sample.
2. Inverse pole gure. IPFs are two-dimensional stereographic plots that use the crystal axes as the reference
frame instead of the sample axes. They show the distribution of crystallographic axes parallel to a specic sample direction. In the case of uniaxial deformation
geometry (extrusion, uniaxial compression, etc.) IPFs
are very convenient for describing the texture, since in
contrast to PFs they require only one well-dened sample direction.
3. Orientation distribution function. The full three-dimensional representation of crystallographic texture is given
by the ODF, which is not measured directly but rather
calculated by combining the data from several experimentally measured pole gures. An ODF describes the
orientation of each crystal relative to three Euler angles,
i.e. u1, U and u2 (Bunge notation) that, in a simplied
sense, dene the dierence in orientation between the
crystallographic axes and the sample axes. The threedimensional coordinate system, whose axes are given

839

by the three Euler angles, is known as the orientation


space or Euler space. Although many advanced softwares nowadays can easily represent ODFs in threedimensional Euler space, it is still more convenient to
use a two-dimensional representation by means of sections through the Euler space (commonly along u2).
The maximum size of Euler space is dened by u1
and u2 ranging from 0 to 360 and U ranging from
0 to 180. In many cases (e.g. conventional rolling) a
reduced Euler space is sucient to describe an
orientation.

A.1. Analysis of sheet textures in Mg


Fig. 18 shows an example of a sheet texture of pure magnesium after hot rolling at 400 C up to 75% thickness
reduction and subsequent annealing at 300 C for 1 h, represented in terms of the (0 0 0 2) and h1 0 
1 0g PFs
(Fig. 18b), the IPF relative to the sheet normal ND
(Fig. 18c), and ODF sections at u2 = 0 and u2 = 30.
The (0 0 0 2) pole gure (also known as the basal pole gure) represents the orientation of the [0 0 0 1] axis, i.e. c-axis,
with respect to the sheet directions ND, RD and TD. It can
be seen that the highest pole density is located in the center
of the pole gure, which corresponds to the c-axis of the
hexagonal crystals being aligned with ND (cf. schematic
in Fig. 18a). A maximum intensity of 14 indicates that
the pole density at that ideal position (ND) is 14 times random. Other positions tilted away from ND have less intensity and correspond to the texture scatter component. The
basal pole gure in Fig. 19b shows a maximum basal pole
scatter of 32 about ND.
While the (0 0 0 2) pole gure is useful in describing the
rolling texture (termed the basal texture), it gives no further information if the texture has a brous nature or if it
has a certain component that is preferably aligned with
RD. For this reason, a second pole gure is needed for a
more complete description of the texture. The f1 0 
1 0g
pole gure (also known as the prismatic pole gure) reveals
that the 1 0 1 0 axes are preferentially aligned with the
sheet transverse direction TD. Due to the sixfold symmetry
of basal plane, this texture component appear six times in
the f1 0 1 0g pole gure, i.e. located at TD and at 30
from RD (cf. schematic in Fig. 18a). In terms of the IPF,
Fig. 18c analogously reveals that the sheet texture is
characterized by pronounced c-axis alignment with ND.
Additional information on the alignment of the h1 1 
2 0i
axes with RD requires a second IPF with respect to RD
or TD.
The orientations in Figs. 18d and 19 (ideal ODF positions for important texture components in Mg) are represented by the three Euler angles (u1 U u2) and their
corresponding MillerBravais indices given by {h k i l}
hu v t wi, which means that the {h k i l} planes of these grains
lie parallel to the sheet plane, whereas their hu v t wi
directions point parallel to the rolling direction. The

840

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843

Fig. 18. Example of a pure magnesium sheet texture obtained after hot rolling (400 C/75% reduction) and annealing at 300 C for 1 h represented
schematically (a), by means of (b) recalculated (0 0 0 2) and f1 0 1 0g pole gures, (c) inverse pole gure relative to ND, (d) ODF sections at u2 = 0 and
u2 = 30.

representation is restricted to (0 6 u1 6 90, 0 6 U 6 90,


0 6 u 2 6 60). The orientation (30 0 0) or
(0 0 0 1)h1 1 
2 0i (at u2 = 0) indicates that the (0 0 0 1)
planes, i.e. basal planes, lie parallel to the sheet plane
and the h1 1 
2 0i directions are aligned parallel to the rolling direction. For an ODF section at u2 = 30, the same
orientation shows at u1 = 0 and 60 and U = 0. With

increasing U the basal plane tilts away from ND


(Fig. 19). The orientation (0 20 0) describes a tilt of basal
plane by 20 from ND toward TD, whereas the orientation (90 20 0) describes a tilt of basal plane by 20
from ND toward RD. For a more comprehensive review
of texture analysis in hexagonal materials, the reader is
referred to Ref. [89].

Fig. 19. Position of dierent ideal orientations in Euler space represented in terms of ODF section at u2 = 0.

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843

841

Fig. 20. fcc rolling texture: a- and b-bers.

Fig. 21. Typical fcc orientations and their rotations in Euler space.

A.2. Analysis of sheet textures in aluminum


The Cube texture as the main recrystallization texture of
aluminum sheet, serves as a very good example to illustrate
the complex methods of texture presentation since here all
axes of the crystallographic (cubic) lattice are in coincidence with the sample coordinates (sheet normal ND, rolling direction RD and transverse direction TD). In the ODF
(Fig. 13a and b it simply occupies each corner of the Euler
space (all Euler angles are 0 or 90), while in the commonly used {1 1 1} pole gure (Fig. 13c) it simply displays
the four {1 1 1} planes (= cell diagonals), which are

strongly detected rst in an X-ray diraction experiment.


All orientations present in the {1 1 1} or {200} pole gures
and in the Euler space can be found on the interactive elearning website AluMatter [90].
The rolling texture of aluminiumlike all fcc metals and
alloysconsists mainly of four (ideal) orientations situated
on two principal rolling texture bers, called a- and b-ber.
The a-ber represents all crystals oriented with a h1 1 0i
direction parallel to the normal direction, ND, the b-ber
can best be approximated by a h1 1 0i axis tilted 60 from
ND towards RD. The ideal orientations are plotted in
Fig. 20a (with all variants) in a {1 1 1} pole gure and the

842

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843

commonly used names and (low-digit) Miller Indices. Since


information is lost when using two-dimensional pole gure
data due to overlapping of the poles (Fig. 20a), the presentation in form of three-dimensional ODFs is necessary.
Fig. 20b shows ideal die orientations in three-dimensional
orientation space (Euler space in Bunge notation [87]). A
comprehensive overview on rolling texture description
and evolution for various fcc metals is given in Ref. [77].
Fig. 21 shows two important sections of the Euler space,
with characteristic orientations and their systematic rotations and other related bers. These can best be observed
in u2 = 0 and u2 = 45 sections, which are therefore used
to describe many texture eects in three dimensions, as well
as recrystallization textures (e.g. PSN textures in Fig. 14).
An introduction to basic texture analysis, including macroand microtextures, and modern orientation mapping can
be found in Ref. [91].
A.3. Mechanical anisotropy
The properties of most metals are never completely uniform in all directions and some degree of anisotropy is
always present. For sheet metal forming (e.g. deep drawing) the average Lankford value r and the planar anisotropy Dr are important parameters. The r-values of rolled
sheet are usually calculated from the true plastic strain
ratios taking the volume as constant during plastic
deformation:
ew
ew
r
;
A1
et
el ew
where ew, et and el are the true strains along the specimen
width, thickness and length directions, respectively (e.g.
calculated at a tensile strain of 0.1). In a textured material
(not only in hexagonal metal sheets, but also in cubic metals), the r-value is highly anisotropic. The average r-value
and the planar anisotropy Dr (which inuences the strain
distribution in the plane of the sheet) are commonly dened as:
r rRD 2r45 rTD =4;

A2

Dr rRD rTD  2r45 =2:

A3

To simulate plastic anisotropy from texture data a suitable viscoplastic self-consistent (VPSC) polycrystal plasticity code has been developed [76] and is frequently used.
References
[1] Avedesian M, Baker H, editors. ASM specialty handbook, magnesium and magnesium alloys. Materials Park, OH: ASM International;
1999.
[2] Jaschik T, Haferkamp H, Niemeyer M. New magnesium wrought
alloys. In: Kainer KU, editor. Magnesium alloys and their applications. Weinheim: Wiley-VCH; 2000.
[3] Blawert C, Fechner D, Hoche D, Heitmann V, Dietzel W, Kainer
KU, et al. Corros Sci 2010;52:2452.
[4] Von Mises R. Z Angew Math Mech 1928;8:161.
[5] Obara CT, Yoshinaga H, Morozumi S. Acta Metall 1973;21:845.

[6] Hardie D, Parkins RN. Philos Mag 1959;4:815.


[7] Massalski TB. Binary alloy phase diagrams. Materials Park,
OH: ASM International; 1992.
[8] Agnew SR, Horton JA, Yoo MH. Metall Mater Trans A 2002;33:851.
[9] Al-Samman T. Acta Mater 2009;57:2229.
[10] Agnew SR, Horton JA, Lillo TM, Brown DW. Scripta Mater
2004;50:377.
[11] Mukai T, Yamanoi M, Watanabe H, Higashi K. Scripta Mater
2001;45:89.
[12] Kaia M, Horitaa Z, Langdon TG. Mater Sci Eng A 2008;488:117.
[13] Al-Samman T, Gottstein G. Mater Sci Eng A 2008;490:411.
[14] Hielscher R, Schaeben H. J Appl Crystallogr 2008;41:1024.
[15] Al-Samman T, Li X, Ghosh Chowdhury S. Mater Sci Eng A
2010;527:3450.
[16] Sakai T, Jonas JJ. Acta Metall 1984;32:189.
[17] Humphreys FJ, Hatherly M. Recrystallization and related annealing
phenomena. Oxford: Pergamon Press; 1995.
[18] Hutchinson WB. Acta Metall 1989;37:1047.
[19] Sellars CM. Modeling of structural evolution during hot rolling
processes. In: Hansen N, Juul Jensen D, Leers T, Ralph B, editors.
Annealing processrecovery, recrystallization and grain growth.
Roskilde: GH-Tryk ApS Odense; 1986.
[20] Ponge D, Gottstein G. Acta Metall 1998;46:69.
[21] Barnett MR, Sullivan A, Stanford N, Ross N, Beer A. Scripta Mater
2010;63:721.
[22] Ion SE, Humphreys FJ, White SH. Acta Metall 1982;30:1909.
[23] Gorkaya T, Molodov KD, Molodov DA, Gottstein G. Acta Mater
2011;59:5674.
[24] Backx P. Improved formability of magnesium and magnesium alloy
AZ31 by texture and microstructure control. PhD thesis, Ghent
University; 2006.
[25] Li X, Jiao F, Al-Samman T, Ghosh Chowdhury S. Scripta Mater
2012;66:159.
[26] Rokhlin LL. Magnesium alloys containing rare earth metals. London
and New York: Taylor & Francis; 2003.
[27] Ball EA, Prangnell PB. Scripta Metall Mater 1994;31:111.
[28] Bohlen J, Nurnberg MR, Senn JW, Letzig D, Agnew SR. Acta Mater
2007;55:2101.
[29] Yi S, Bohlen J, Heinemann F, Letzig D. Acta Mater 2010;58:592.
[30] Klar R, Lucke K. Z Metall 1968;59:194.
[31] Nadella RK, Samajdar I, Gottstein G. Static recrystallization and
textural changes in warm rolled pure magnesium. In: Kainer KU,
editor. Magnesium alloys and their applications. Weinheim: WileyVCH; 2003.
[32] Gottstein G, Al-Samman T. Mater Sci Forum 2005;495497:623.
[33] Wagner F, Bozzolo N, Van Landuyt O, Grosdidier T. Acta Mater
2002;50:1245.
[34] Molodov DA, Bozzolo N. Acta Mater 2010;58:3568.
[35] Stanford N, Barnett M. Scripta Mater 2008;58:179.
[36] Huppmann M, Gall S, Muller S, Reimers W. Mater Sci Eng A
2010;528:342.
[37] Robson JD, Twier AM, Lorimer GW, Rogers P. Mater Sci Eng A
2011;528:7247.
[38] Grobner J, Schmid-Fetzer R. Scripta Mater 2010;63:674.
[39] Stanford N. Mater Sci Eng A 2010;527:2669.
[40] Bohlen J, Yi S, Letzig D, Kainer KU. Mater Sci Eng A
2010;527:7092.
[41] Stanford N, Barnett M. Mater Sci Eng A 2008;496:179.
[42] Stanford N, Sha G, Xia JH, Ringer SP, Barnett MR. Scripta Mater
2011;65:919.
[43] Gottstein G. Physical foundations of materials science. 1st ed. Berlin: Springer Verlag; 2004.
[44] Li X, Al-Samman T, Mu S, Gottstein G. Mater Sci Eng A
2011;528:7915.
[45] Al-Samman T, Li X. Mater Sci Eng A 2011;528:3809.
[46] Mu S, Al-Samman T, Mohles V, Gottstein G. Acta Mater
2011;59:6938.
[47] Kelly EW, Hosford WF. Trans AIME 1968;242:5.

J. Hirsch, T. Al-Samman / Acta Materialia 61 (2013) 818843


[48]
[49]
[50]
[51]
[52]
[53]
[54]
[55]
[56]
[57]
[58]
[59]
[60]
[61]
[62]
[63]

[64]
[65]

[66]
[67]

[68]
[69]

[70]

Akhtar A, Teghtsoonian E. Acta Metall 1969;17:1339.


Akhtar A, Teghtsoonian E. Acta Metall 1969;17:1351.
Caceres CH, Rovera DM. J Light Metals 2001;1:151.
Raeisinia B, Agnew SR. Scripta Mater 2010;63:731.
Stanford N, Atwell D, Barnett M. Acta Mater 2010;58:6773.
Couling SL, Pashak JF, Sturkey L. Trans ASM 1959;51:94.
Chino Y, Kado M, Mabuchi M. Acta Mater 2008;56:387.
Sandlobes S, Zaeerer S, Schestakow I, Yi S, Gonzales-Martinez R.
Acta Mater 2011;59:429.
Joliet H. Aluminium die ersten hundert Jahre. Dusseldorf: VDI
Verlag; 1988.
Bartz WJ. Aluminum materials technology for automobile construction. London: Mech Eng Publ; 1993.
Hirsch J. Mater Sci Forum 1997;242:33.
Bloeck M, Timm J. Aluminum 1994;70:87.
Falkenstein HP, Gruhl W, Scharf G. Metall Tech 1983;37:1197.
Phillips VA. J Inst Metals 1952;81:601.
Engler O, Lochte L, Hirsch J. Acta Mater 2007;55:5449.
Hirsch J. Introduction. In: Hirsch J, editor. Virtual fabrication of
aluminum products: microstructural modeling in industrial aluminum
production. Weinheim: Wiley-VCH Verlag; 2006.
Hirsch J. Textures in industrial processes and products. In: Tewari A,
et al., editors. Textures of materials. Zurich: TransTech; 2011.
Banabic D, Erman Tekkaya A. Numerical simulation and material
models of aluminium sheet forming. In: Hirsch J, editor. Virtual
fabrication of aluminum products: microstructural modeling in industrial aluminum production. Weinheim: Wiley-VCH Verlag; 2006.
Barlat F, Lian J. Int J Plast 1989;5:51.
Hirsch J. AlMn1Mg1 for beverage cans. In: Hirsch J, editor. Virtual
fabrication of aluminum products: microstructural modeling in industrial aluminum production. Weinheim: Wiley-VCH Verlag; 2006.
Uno T, Baba Y. Aluminum 1987;63:1243.
Brunger E, Engler O, Hirsch J. AlMgSi sheet alloys for autobody
applications. In: Hirsch J, editor. Virtual fabrication of aluminum
products: microstructural modeling in industrial aluminum production. Weinheim: Wiley-VCH Verlag; 2006.
Liang Z, Chang CST, Abromeit Ch, Banhart J, Hirsch J. The eect of
Cu and Cr on clustering and precipitation in AlMgSi alloys. In:

[71]

[72]

[73]
[74]
[75]
[76]
[77]
[78]
[79]
[80]
[81]
[82]

[83]
[84]

[85]
[86]
[87]
[88]
[89]
[90]
[91]

843

Weiland H, et al., editors. Proc. 13th international conference on


aluminum alloys. TMS and Wiley; 2012.
Hirsch J. In: Lumley R, editor. Fundamentals of aluminum metallurgy: production, processing and applications. Cambridge: Woodhead Publishing; 2010.
Kalu PN, Humphreys FJ. Deformation mechanisms in two-phase Al
alloys. In: Sheppard T, editor. Al technology 86. London: Institute of
Metals; 1986.
Engler O, Hirsch J. Mater Sci Eng A 2002;336:249.
Engler O, Hirsch J. Int J Mater Res 2009;100:564.
Lucke K, Darmann C, Hirsch J. Trans Ind Inst Metals 1985;38:496.
Lebensohn RA, Tome CN. Acta Metall Mater 1993;41:2611.
Hirsch J, Lucke K. Acta Metall 1988;36:2863.
Hirsch J. In: Chandra T, editor. Proc. recrystallization 90. Australia: Wollongong; 1990.
Hirsch J, Engler O. In: 16th RIS int symp materials science; 1995.
Engler O, Hirsch J. Mater Sci Forum 1996;217222:479.
Hirsch J, Lucke K. Acta Metall 1985;33:1927.
Hirsch J, Lucke K. Al-alloys: their physical and mechanical
properties. In: Starke EA, Sanders TH, editors. Proc. 1st international conference on aluminum alloys. London: Chameleon Press;
1986.
Engler O, Heckelmann I, Rickert T, Hirsch J, Lucke K. Mater Sci
Technol 1994;10:771.
Engler O, Yang P, Gottstein G, In Starke EA, Sanders TH, editors.
Proc. 4th international conference on aluminum alloys. Atlanta, GA;
1994.
Toth LS, Hirsch J, Van Houtte P. Int J Mech Sci 1996;38:1117.
Brinkman HJ, Engler O, Hirsch J, Schroder D. Aluminium Int Today
2011;23:41.
Bunge HJ. Texture analysis in materials science. London: Butterworths; 1982.
Hirsch J, Lucke K. Text Microstruct 1988;89:131.
Wang YN, Huanga JC. Mater Chem Phys 2003;81:11.
AluMatter. <http://aluminium.matter.org.uk> [anisotropy/texture
representation].
Randle V, Engler O. Introduction to texture analysis. Amsterdam: Gordon & Breach; 2000.

You might also like