Sei sulla pagina 1di 12

Journal of Light Metals 1 (2001) 6172

www.elsevier.com/locate/ligandmet

Development of the as-cast microstructure in magnesiumaluminium


alloys
Arne K. Dahle *, Young C. Lee, Mark D. Nave, Paul L. Schaer, David H. StJohn
?tul=0>CRC for Cast Metals Manufacturing (CAST), Department of Mining, Minerals and Materials Engineering, The University of Queensland,
Brisbane Qld 4072, Australia

Abstract
This paper presents an overview of several projects undertaken at CAST to increase our understanding of the solidication
characteristics of MgAl alloys. With the increased use of magnesium alloys, and with casting dominating as a production route,
there is a need for a more comprehensive understanding of the mechanisms of solidication and defect formation to allow further
optimisation of alloys and casting processes. The paper starts with considering the formation of the primary magnesium dendrites
and the means for grain renement of magnesiumaluminium alloys. The MgAl system is then shown to display a range of eutectic
morphologies for increasing aluminium content, ranging from a divorced structure, through several intermediate structures, to a
fully lamellar structure at the eutectic composition. The eutectic also inuences discontinuous precipitation which occurs in the
aluminium-rich regions of the magnesium phase. The paper concludes with a section on porosity formation as a function of aluminium content and an outline of the mechanism responsible for the formation of banded defects in magnesium alloys, particularly
in products made in pressure assisted casting processes. 2001 Elsevier Science Ltd. All rights reserved.
Keywords: Magnesium; Solidication; Casting defects; Eutectic growth; Grain renement; Porosity

1. Introduction
Most commercial magnesium alloys are based on the
magnesiumaluminium system and casting is currently
the most commonly used production process for magnesium components [1]. Among the alloys used, AZ91,
AM60 and, to a lesser extent, AM50 dominate. The A in
the alloy designation indicates that aluminium is the
main alloying element and the rst numeral is the approximate concentration of aluminium in wt%. AZ91
therefore contains 9 wt% aluminium, and the Z indicates
it also contains about 1 wt% zinc (actually 0.7 wt% Zn).
The range of aluminium contents for the commercial
alloys is 39 wt% Al, from AZ31, a wrought alloy
composition cast as billet, to AZ91.
The MgAl alloys are relatively cheap compared with
other magnesium alloys available. They are readily castable, particularly by high-pressure die casting, and
exhibit good mechanical properties [1]. An important
feature of these alloys is that they can be cast into long
and thin sections by high-pressure die-casting. Typical

Corresponding author.
E-mail address: A.Dahle@minmet.uq.edu.au (A.K. Dahle).

applications that utilise this feature include instrument


panels, steering wheels and seat frames [1]. Although
magnesium alloys containing aluminium generally possess good mechanical properties, ternary alloys with
zinc, manganese, silicon and rare-earth elements in addition to aluminium are used to obtain improved mechanical properties. Zinc is added to improve the room
temperature strength and uidity while silicon is added
to improve the creep strength of the alloys by forming
Mg2 Si particles on the grain boundaries [2]. Addition of
manganese is required to control the corrosion behaviour, and magnesium alloys with aluminium and
manganese (AM60, AM50) are commonly used for
components where good ductility and impact strength
are required.
Fig. 1 shows the magnesiumaluminium equilibrium
phase diagram. The maximum solid solubility at the
eutectic temperature is about 13 wt% Al, and a eutectic
between Mg and the intermetallic Mg17 Al12 appears at
about 33 wt% Al. All the aluminium contents used in the
commericial alloys are below the maximum solid solubility limit and the alloys therefore solidify with a
primary a-magnesium phase. The equilibrium microstructure for all the alloys is 100% a-magnesium, but
non-equilibrium, metastable, eutectic normally forms

1471-5317/01/$ - see front matter 2001 Elsevier Science Ltd. All rights reserved.
PII: S 1 4 7 1 - 5 3 1 7 ( 0 0 ) 0 0 0 0 7 - 9

62

A.K. Dahle et al. / Journal of Light Metals 1 (2001) 6172

during solidication and is present in the as-cast


microstructure in MgAl alloys down to about 2 wt%
Al.
Since solidication of MgAl alloys results in a
structure consisting of primary dendrites and eutectic, it
is tempting to make comparisons between these alloys
and the other major commercial light casting alloys, the
AlSi alloys. However, there are some very signicant
dierences. First, the eutectic is metastable in the MgAl
alloys, as discussed above. Heat treatment can therefore
result in a complete dissolution of the Mg17 Al12 intermetallic phase. In contrast, the Si phase in AlSi alloys is
reasonably stable during heat treatment and therefore
remains after heat treatment, although it spheroidises
resulting in some improvement in elongation. Second,
the eutectic microstructure in AlSi alloys is an irregular
eutectic with Si as the faceted phase and the coarse Si
morphology is often modied to reduce the size and
rene the eutectic to improve the mechanical properties.
In MgAl alloys with more than about 20 wt% Al,
the eutectic is a regular eutectic where both phases are
non-faceted [4,5]. However, the volume fraction of
eutectic decreases as the Al-content is decreased, and
the eutectic morphology gradually transforms to a divorced eutectic [4,5]. Another dierence is in the volume
fraction of eutectic, which is much smaller in the MgAl
alloys compared to that in AlSi foundry alloys.
MgAl alloys display a wide freezing range, Fig. 1,
and these alloys are therefore susceptible to a range of
casting defects including segregation, porosity and hot
tearing. A distinctive kind of defect often forms when,
and because, the alloys are used to cast long, thin sections by high-pressure die casting [68]. This defect has
the form of bands of segregation, porosity and tears that
usually run parallel to the surface of the casting. Although it has been shown that these defects may form in

AlSi alloys in pressure assisted casting processes [8],


these defects are much more common in MgAl alloys
and they are now regarded as a major impediment to
obtaining good mechanical performance from cast
components [9].
Magnesium components are most commonly produced by high-pressure die casting and, sand and lowpressure die casting are only used to a small extent.
Sand-casting is used to some extent for the production
of magnesium alloys not containing aluminium for
aeronautical components [1]. There are, however, some
components, such as wheels, where a low-pressure or
gravity die casting process may be more suitable. A
problem is that no reliable grain rener addition exists
for MgAl alloys and, thus, the benets of a ne grain
size on the mechanical properties cannot be reliably
achieved [1].
Also, alloy optimisation to minimise porosity has not
been undertaken to the degree that it has been for AlSi
alloys. A better understanding of how defects form
during solidication is required to be able to obtain the
best performance from cast MgAl alloys.
This paper presents an overview of a range of projects
that has been undertaken to improve our understanding
of the solidication characteristics of MgAl alloys. It
follows the solidication process of the alloys, beginning
with the nucleation, grain renement and growth of the
primary a-magnesium phase. Next, the formation of the
eutectic is considered showing a range of morphologies
of the eutectic Mg17 Al12 phase. The last section considers defect formation in MgAl alloys, rst focussing on
the porosity characteristics as a function of aluminium
content and, nally, the mechanism responsible for the
formation of shear defects and porosity in pressure assisted casting processes commonly used for production
of MgAl components.
2. Nucleation and growth of -magnesium equiaxed dendrites

Fig. 1. MgAl equilibrium phase diagram, adapted from [3].

The solidication sequence of MgAl alloys starts


with nucleation of primary magnesium (a-Mg) in the
temperature range 650600C, ranging from the melting
point of pure magnesium to the liquidus temperature of
Mg 9 wt% Al, covering the aluminium contents used in
most commercial alloys. Later solidication reactions
involve the formation of eutectic phases, with the Mg
Mg17 Al12 eutectic reaction occurring at 437C. A typical
microstructure of MgAl alloys is shown in Fig. 2,
where well-developed primary a-Mg dendrites with
secondary arms (A) showing sixfold symmetry are
clearly visible. The eutectic of Mg17 Al12 (B) and a-Mg
solid solution (C) is located in the interdendritic regions.
According to the MgAl equilibrium phase diagram,
Fig. 1, the eutectic phase (Mg17 Al12 ) is expected to ap-

A.K. Dahle et al. / Journal of Light Metals 1 (2001) 6172

Fig. 2. Micrograph of fully developed dendrites in a Mg15 wt% Al


alloy permanent mould casting [3]. The microstructure of the alloy is
similar to 9 wt% aluminium alloy, but the dendritic structure is more
clearly visible than normally observed in a 9 wt% aluminium alloy. The
magnesium dendrites have a characteristic sixfold symmetric shape
(A). The white phase between the dendrites is secondary eutectic phase
Mg17 Al12 (B) and the dark regions between the dendrites are aluminum-rich solid solution (C). Some of the enriched solid solution forms
within the Mg17 Al12 phase (partially divorced eutectic). A more detailed explanation of eutectic phase is presented later in this article.

pear when the aluminium content reaches around 13


wt%. However, the eutectic phase appears in alloys
containing as little as 2 wt% Al for non-equilibrium
cooling conditions normally encountered in castings
[10].
Fig. 3 shows the microstructural changes with increasing aluminium content. A small addition of aluminium to pure magnesium leads to a morphological

63

change of the primary phase from a cellular to a dendritic structure. Rosette-like globular equiaxed grains
form with aluminium-rich solid solution between the
dendrite arms. As the aluminium content is increased
further to 5 wt%, dendrites with pools of eutectic phase
between the dendrite arms start to develop and, when
the aluminium content is further increased, a fully developed dendritic structure with sharp tips is observed.
The addition of small amounts of alloying elements
such as zinc, manganese, silicon and rare-earths to Mg
Al alloys has little eect on nucleation of the primary
phase since these elements are mostly segregated to form
secondary phases well after the primary phase has nucleated [11].
Grain renement is among the important practices
used to improve the properties of castings. It is an essential and fundamental approach since grain size signicantly inuences the mechanical properties of the
castings, and the grain size is usually determined at an
early stage of solidication by nucleation of the dendrites.
Several methods of grain renement have been developed. One of the rst methods was grain renement
by a simple thermal treatment prior to casting, the socalled `superheating treatment'. This method involves
rapid cooling of the melt to the desired casting temperature after a short holding time at an elevated temperature, generally between 150C and 260C above the
equilibrium liquidus temperature of the alloy [12]. Despite the successful grain renement achieved by the
superheating method, alternative techniques were
sought due to several practical problems, mainly related
to the higher operating temperatures involved.
Successful grain renement has been reported by the
addition of ferric chloride (Elnal process) in magne-

Fig. 3. Micrographs of magnesiumaluminum alloys with increasing aluminum content. The transition from a globular dendritic structure to a fully
developed dendritic structure with increasing aluminium content is readily noticeable.

64

A.K. Dahle et al. / Journal of Light Metals 1 (2001) 6172

sium alloys containing aluminium and manganese. The


amount of grain renement achieved by the Elnal
process is somewhat similar to that achieved by superheating [12,13]. However, due to the detrimental eect
on corrosion resistance from the addition of Fe, the
Elnal process has not attracted industrial attention.
The addition of carbon to the melt (carbon inoculation) oers more practical advantages accompanied by
lower operating temperatures and less fading. Various
carbon-containing agents such as organic materials
(C2 Cl6 ; CCl4 ) [12,14], SiC particles [15], or granular
graphite [16] have been reported produce successful
grain renement in magnesium alloys containing aluminium.
A number of hypotheses have been proposed to explain the mechanisms by which carbon inoculation
methods cause grain renement. However, none of them
provides an adequate explanation of the mechanisms
due to a lack of understanding of the fundamental factors involved, such as the identication of the active
nucleants and/or solute elements with strong segregating
tendency. The formation of Fe and/or Mn containing
compounds is considered to cause grain renement in
the superheating and Elnal treatments, while the formation of Al4 C3 particles, regarded to be very potent
nucleation sites for magnesium, is believed to produce
renement in the carbon inoculation methods.
Recent work [1719] on the grain renement of aluminium alloys has shown that grain renement can be
facilitated by two mechanisms. The rst is the formation
of crystals in the thermally undercooled region near the
walls of the mould early in the casting process. These
crystals are then carried into the bulk of the melt by
convection currents. The other mechanism is a result of
constitutional undercooling generated by the growth of
a grain adjacent to a nucleant particle suspended in the
melt [1719]. In both cases there are two factors which
can enhance the number of successful nucleation events.
The rst is the solute elements present in the melt and
the other is the number and potency of the nucleant
particles. Fig. 4 shows the relative eect of these two

Fig. 4. Relative grain size as a function of solute level for a range of


nucleant potencies (after [20]).

factors where it can be seen that when potent nucleant


particles are present, addition of solute causes the grain
size to decrease rapidly at rst and then more slowly at
higher solute contents. As shown below, dierent solute
elements cause this decrease to occur at dierent rates
depending on their segregating power. Fig. 4 also shows
that as the potency of the nucleant particles decreases,
not only is the grain size obtained larger, but the eect of
solute additions on grain size is decreased.
Through a simplication, the eect of the solute can
be dened by the alloy's growth restriction factor when
the potency of the nucleant particles is very high [20].
The
P growth restriction factor (GRF) is dened by
i mi C0;i ki 1 where m is the slope of the liquidus
line, C0 the initial composition, and ki is the equilibrium
partition coecient for element i [21]. A large GRF indicates that the growing crystal generates constitutional
undercooling quickly and the liquid around the adjacent
nucleants is therefore more quickly undercooled suciently to allow a stable nucleus to form on the nucleant
particle compared with an alloy having a small GRF.
The eect of nucleant potency is that a higher potency
(i.e., a lower amount of undercooling required for nucleation) will result in the next nucleation event occurring relatively sooner. Therefore, a high-potency
nucleant combined with a melt composition with a large
growth restriction factor is likely to result in a small
grain size. This is the principle behind the AlTiB grain
reners for aluminium alloys where the TiB2 particles
are the potent nucleant particles and the excess titanium
ensures that the melt composition has a high-growth
restriction factor. Some alloying elements, such as aluminium, calcium, silicon and zirconium, have been reported to produce signicant grain renement in pure
magnesium due to their strong segregating power [22].
The closest to satisfying the conditions of the potent Al
TiB master alloys in magnesium alloys is the use of
zirconium master alloy to rene alloys that do not
contain Al. In this case a very ne grain size can be
obtained. This is because zirconium is the most strongly
segregating element in magnesium [22,23]. The exact
nature of the particles present in a MgZr melt is unknown, but they appear to be very potent. It is worth
noting that grain renement by zirconium addition is
not eective in magnesium alloys that contain aluminium due to an undesirable interaction between aluminium and zirconium.
The above-described mechanism of nucleation of the
primary phase has the implication that if all alloy
compositions are converted to their corresponding
growth restriction factors, then the plots of grain size
versus GRF should be the same for all alloys when the
number and potency of nucleant particles are equal.
Fig. 5 shows that this is approximately true for the binary additions of Si, Ca and Zr. However, the result for
Al is very dierent. By considering Figs. 4 and 5, this

A.K. Dahle et al. / Journal of Light Metals 1 (2001) 6172

Fig. 5. Relationship between grain size and growth restriction factor


(GRF) for binary Mg alloys for the elements Al, Si, Zr and Ca. The
results indicate far less potent nucleant particles in MgAl alloys
compared to other alloys since the saturation level of grain density is
only reached at much higher values of the GRF.

indicates that the nucleant particles existing in MgAl


alloys are much less potent. It has been speculated that
Mn causes grain coarsening in these alloys [24]. Also, it
was recently shown that the grain size of AZ91 varies
substantially depending on the purity of the ingots used
to make the alloy [22,23]. It therefore appears that aluminium in combination with manganese and the impurity elements aects the potency of the nucleants in the
melt.
Grain size behaviour with direct particle additions of
SiC, AIN and Al4 C3 is shown by the macrographs in
Fig. 6. All of these particles are expected to be good
nucleants for magnesium due to a small lattice disregistry towards magnesium. The macrographs also conrm that most of the particles possess a relatively good
nucleation potency. These results may support numerous hypotheses that these particles, particularly Al4 C3 ,
are active nucleants for magnesium in commericial alloys since MgAl alloys can be rened by the addition of
carbonaceous substances.
Although our understanding of the principles of grain
renement in magnesium alloys has improved signicantly, a reliable and readily added master alloy for grain
renement of MgAl alloys still remains to be developed.
Work still remains to determine the eect of Al in
combination with Mn, Fe and other impurity elements
on grain size and the results may hold a key to understanding grain renement in the MgAl alloy system.
3. Eutectic growth
As mentioned before, the cooling rates in commercial
casting processes are generally sucient to cause some

65

eutectic to form during solidication of magnesium alloys containing more than 2 wt% Al [10]. Die-castings of
the common commercial magnesium alloys, AZ91,
AM50 and AM60, therefore contain a signicant volume fraction of eutectic. Understanding eutectic solidication in these alloys is important for two main
reasons. The rst is that this solidication event controls
the size, shape and distribution of the more brittle
b-Mg17 Al12 phase in the nal microstructure, which, in
turn, is likely to inuence both the ductility [10] and
creep strength [2527] of the alloys. The second is that,
being the nal stage in the solidication process, eutectic
growth aects feedability at a crucial stage, when feeding is interdendritic and large pressure dierentials are
required to draw liquid through the dendritic network.
A dierence in eutectic growth mode could have a large
eect on the ease with which liquid can be drawn
through the dendritic network, and therefore on the
formation of porosity in these alloys.
The eutectic exhibits a wide range of morphologies in
hypoeutectic MgAl alloys depending on composition
and cooling rate [4,5]. Alloys with aluminium contents
approaching the eutectic composition (33 wt% Al) tend
to display regular lamellar or brous eutectic microstructures [28], while those with aluminium contents less
than about 10 wt% Al (i.e. commercial alloys) exhibit
eutectic morphologies that are generally referred to as
fully or partially divorced. A fully divorced morphology
(Fig. 7(a)) is where the two eutectic phases are completely separate in the microstructure. Each interdendritic region consists of a single b-Mg17 Al12 particle
surrounded by `eutectic' a-Mg, which has grown from
the primary dendrites. A partially divorced eutectic
morphology (Fig. 7(b)) is characterised by `islands' of
eutectic a-Mg within the b-Mg17 Al12 phase, but the bulk
of the a-Mg is still outside the Mg17 Al12 particle, i.e. the
volume fraction of a-Mg within the Mg17 Al12 particle is
much lower than the proportion predicted by the equilibrium phase diagram.
The eects of aluminium content, zinc content and
cooling rate on eutectic morphology in permanent
mould cast alloys are shown schematically in Fig. 8. The
eutectic tends to become less divorced with increasing
aluminium content [4], but more divorced with increasing zinc content [5] and cooling rate [4,5]. The main
mechanisms by which composition and cooling rate may
aect eutectic morphology are discussed in detail by
Nave et al. [4,5], and are related to the location of the
coupled zone and the undercooling during solidication.
An understanding of these mechanisms points the way
towards methods for modication of the MgMg17 Al12
eutectic. For instance, the addition of a ternary element
which does not partition as strongly as zinc to the liquid
during a-Mg growth, and to the Mg17 Al12 phase during
eutectic growth, is likely to cause a less divorced eutectic
morphology than the addition of an equivalent amount

66

A.K. Dahle et al. / Journal of Light Metals 1 (2001) 6172

Fig. 6. Macrographs of (a) Mg1Al, (b) Mg1Al+1vol% AlN-particles, (c) Mg1Al+1vol% Al4 C3 -particles, and (d) Mg1Al+1vol% SiC-particles.

Fig. 7. Fully divorced (a) and partially divorced (b) eutectic morphologies in a hypoeutectic MgAl alloy. The lightest areas are b-Mg17 Al12 and the
darkest area are high Al content (`eutectic') a-Mg. The grey areas are primary a-Mg dendrites, showing coring from the low Al content areas near the
centres of their arms (light grey) to the higher Al content areas near the edges of their arms (dark grey).

A.K. Dahle et al. / Journal of Light Metals 1 (2001) 6172

Fig. 8. The eect of aluminium content, zinc content and cooling rate
on eutectic morphology in permanent mould cast hypoeutectic MgAl
alloys.

of zinc, particularly if it also increases the distance between the Mg17 Al12 phase boundary and the eutectic
point.
In terms of feedability during the critical last stages of
solidication, both eutectic morphology and its mechanism of formation are important. Independent nucleation and growth of the b-Mg17 Al12 phase in the
interdendritic liquid are likely to signicantly increase
the surface area to volume ratio of the feeding channels,
providing a much greater resistance to feeding. On the
other hand, nucleation of the b-Mg17 Al12 phase on
the a-Mg and subsequent growth towards the centre of
the interdendritic channels should allow the feeding
paths to remain more open for longer and tend to promote a sounder casting. Specimens quenched during
directional solidication show quite a good wetting between b-Mg17 Al12 and a-Mg phases and growth appears
to occur from the surface of the dendrite into the interdendritic liquid see Fig. 9.
Feeding during eutectic solidication is also likely to
be aected by the solidication range of the eutectic and
the smoothness of the solid/liquid interface during eutectic growth (i.e. if any phase is a leading phase,
growing ahead of the other phase during coupled
growth), or how isothermal the eutectic growth interface
is. An isothermal, smooth, interface allows easier feeding, while an interface in which one of the phases grows
with a considerable lead over the other phase would
require feeding along increasingly narrow and convoluted paths in order to avoid porosity formation. Zinc,
which is present in the common commercial alloy AZ91,
segregates strongly to the Mg17 Al12 phase during eutectic growth [5], increasing the lead of this phase over
the a-Mg phase and causing the eutectic to solidify with
a less isothermal interface. The addition of zinc could
therefore be expected to promote porosity formation

67

Fig. 9. A section of a directionally solidied Mg9.1Al0.4Zn alloy


that has been quenched during eutectic growth. There appears to be
good wetting between the a-Mg and b-Mg17 Al12 phases.

based on the above considerations. While there is some


evidence that the addition of 2 wt% Zn increases microporosity in sand-cast magnesium alloys containing 2, 4,
8 and 10 wt% Al [29], more detailed work is required to
establish the eect of zinc on porosity formation in Mg
Al alloys.

4. Precipitation reactions
Completion of eutectic solidication does not necessarily mark the end of phase transformations in a cast
magnesiumaluminium alloy. When the cooling rate of
the casting is suciently slow (typical of sand-casting),
precipitation may occur in the supersaturated areas of
the a-Mg. This precipitation may take two forms, continuous (Fig. 10(a)) and discontinuous (Fig. 10(b)) precipitation. The most obvious form, and the form by
which the bulk of the precipitation occurs, is discontinuous precipitation. This involves the growth of
lamellar precipitates of Mg17 Al12 into the a-Mg grains in
a similar manner to the way pearlite colonies grow into
austenite grains during the cooling of steel. The aluminium partitions to the Mg17 Al12 lamellae as they
grow, leaving the a-Mg between the lamellae much
leaner in aluminium than before discontinuous precipitation commenced. The discontinuous precipitation
appears to grow from near the eutectic Mg17 Al12 into the
a-Mg grains, but whether the precipitates actually have
the same orientation as the Mg17 Al12 phase, or whether
they nucleate separately in the supersaturated a-Mg
phase (e.g. on a dislocation) has not been conrmed.
Discontinuous precipitation occurs mostly in the a-Mg
regions near the Mg17 Al12 phase, since these regions
have higher aluminium contents (approx. 1013 wt% Al)

68

A.K. Dahle et al. / Journal of Light Metals 1 (2001) 6172

Fig. 10. (a) Continuous and (b) discontinuous precipitation in alloy AZ91E.

than the centres of the dendrites, where the aluminium


concentrations may be as low as 2 wt% Al.

5. Defect formation
5.1. Porosity characteristics
The solidication sequence of MgAl alloys plays an
important role in the formation of defects. The causes of
microporosity in MgAl alloys has been a contentious
issue since early work conducted by Baker [30] in the
1940s and much debate has occurred concerning which
of the variables, solidication shrinkage or dissolved
gas, contribute to the formation of microporosity.
Initial investigations led to the belief that the formation of microporosity was not signicantly aected by
dissolved hydrogen. The dierence in solubility of hydrogen between the solid and liquid phase is relatively
small compared to aluminium alloys and it was believed
that solidication shrinkage was responsible for porosity
formation. However, this theory was soon dispelled and
investigations measuring the hydrogen content of the
melt before solidication have suggested that dissolved
hydrogen does contribute to the incidence of microporosity [3032].
It is unlikely that the formation of microporosity in
MgAl alloys is isolated to one of the two mechanisms
outlined above. Moreover, solidication shrinkage and
evolution of dissolved gas occur in unison and act collaboratively to form microporosity just as in other alloy
systems [33]. Magnesium alloys solidify relatively slower
than aluminium alloys because of their low thermal
conductivity. Therefore progressive feeding is dicult
and interdendritic feeding is a very important stage
during solidication of magnesium alloys as a result of
the wide freezing range [30]. vrelid et al. [34] reported
that increased aluminium content decreases the solubility of hydrogen in the liquid.

Recent work has shown that the aluminium content


of MgAl alloys strongly aects the volume of porosity
[35]. Cylindrical, unfed, castings were produced for
dierent aluminium contents in the hypoeutectic MgAl
region. The results are reproduced in Fig. 11. The results
show that the peak in porosity occurs at about 9 wt%
Al, slightly below the maximum equilibrium freezing
range as also plotted in the gure. This result can be
compared to a reported maximum in hot tearing susceptibility at about 1 wt% Al in the MgAl system [36].
Maximum hot tearing is normally related to the occurrence of the rst eutectic liquid and a maximum, nonequilibrium, freezing range. Furthermore Campbell [37]
has argued that there is a maximum of hydrostatic
pressure when the rst eutectic liquid appears and that
this point also should correlate with a maximum in
porosity. It is therefore clear from Fig. 11 that porosity
in MgAl alloys does not follow the same trend and
therefore requires a dierent explanation. The low levels
of porosity in pure magnesium and the eutectic alloy (33
wt% Al) are related to the isothermal, or near-isother-

Fig. 11. Percentage porosity versus aluminum content and freezing


range versus aluminium content. Note that the large freezing range
correlates with a greater incidence of porosity.

A.K. Dahle et al. / Journal of Light Metals 1 (2001) 6172

mal, solidication of these alloys. In the other alloys, the


volume fraction of eutectic will increase with increased
aluminium content, which can be expected to reduce
porosity when the volume fraction is sucient. The
reason for the peak in porosity at about 9 wt% must
therefore be related to the worst combination of mushy
zone size, interdendritic feeding, permeability and eutectic volume fraction. When these alloys solidify there
is insucient liquid to feed the shrinkage of the eutectic,
and porosity therefore initially increase with eutectic
volume fraction. The equilibrium solidication range for
the alloys is also indicated in the diagram for the sake of
comparison and if can be observed that the maximum
amount of porosity is located closer to the maximum
equilibrium solidication range rather than the maximum for non-equilibrium conditions predicted to be
around 12 wt% Al. As discussed in the section on eutectic growth there are signicant changes in the eutectic
morphology with increased aluminium content that
certainly may inuence porosity formation and interdendritic ow. Further discussion is provided in [35].
This is an area of research that deserves further study.
5.2. Defect formation in pressure assisted casting processes
It has recently been documented that casting of long
thin sections of magnesium alloys in pressurised casting
processes can result in long, continuous, bands of defects with an outline that follows the contour of the
casting [68]. The appearance of the defect varies from
being highly segregated, through porous, to torn. Fig. 12

69

shows an example of a shear band in an AM60 alloy


casting where the band contains segregated eutectic and
porosity.
A model has been developed which can predict the
occurrence, location and appearance of the bands based
on understanding the solidication characteristics and
the mushy zone mechanical properties [7,8]. This model
can be simplied to describe the mushy material as a
series of thermal contours that divides the solid fraction
and the mechanical behaviour of the mush into four
regions. At an early stage of lling the thermal contours
extend from a temperature near the liquidus temperature
in the centre of the casting to near or below the solidus
temperature at the walls of the die cavity. The top
temperature is dened by the thermal characteristics of
the shot sleeve of the casting machine. Often the liquid
entering the die cavity will have a solid fraction of about
0.2 due to premature solidication in the shot sleeve [38].
The contours are delineated by critical temperatures that
dene a signicant change in the mechanical response to
shear stresses generated during mould lling. The critical
temperatures are the liquidus temperature, the coherency point where the dendrites rst touch each other
causing the rst measured resistance to motion, the
maximum packing point where the dendrites become
fully interlocked, the eutectic temperature where the
microstructure becomes more rigid due to eutectic
bridging joining the dendrites together, and the solidus
temperature [7,39,40]. Table 1 summarises the dierences in mechanical behaviour of the regions between
these contours. During solidication the range of
properties of the mush in the die cavity can be dened by

Fig. 12. (a) A defect band containing segregation of eutectic and porosity in a circular section of a high-pressure die cast AM60. The ow direction is
into the page. Note the very ne grain size on the outside of the defect band compared with the coarser microstructure inside. Higher magnication in
(b) with the region outside the band on the left hand side of the micrograph. (Courtesy: A. Bowles, CAST.)

70

A.K. Dahle et al. / Journal of Light Metals 1 (2001) 6172

Table 1
A summary of behaviour of the partially solidied casting alloy within the zones dened by the critical thermal contoursa

Thermal zone

Microstructure

Feeding mechanisms

Mechanical behaviour

Tl Tch

Dendrites surrounded by liquid

Mass feeding

Tch Tpk

Dendrites touching each other with


liquid lms

Mass feeding with an increasing


proportion of interdendritic feeding
as solid fraction increases

Tpk Te

Dendrites packed together and


mechanically interlocked

Interdendritic and burst feeding

Te Ts

Dendrites surrounded by partially


solidied eutectic

Interdendritic and solid feeding while


eutectic solidication

No strength and liquid-like


behaviour
Low strength with slurry-like or
thixotropic behaviour. Packing
occurs. Shear causes liquid to
segregate to shear plane
Strength increases at a faster rate.
Solid-like behaviour with dendrites
deforming and network fracture
Rapid increase in strength to Ts .
Behaves as a solid

The liquidus Tl , coherency Tch , maximum packing, Tpk , eutectic Te , and the solidus Ts .

the locations of these critical contour boundaries. Additionally, the feeding mechanism within each of these
zones can be estimated (also listed in Table 1). Fig. 13 is
a representation of the change in shear strength with
solid fraction. The values of coherency and maximum
packing solid fractions depend on the dendrite size and
morphology, and small spherical dendrites have larger
values than large irregular dendrites [7,39,40].
When signicant ow has to occur while partially
solidied mush is present, the mush will deform, and the
low-strength regions are the most likely to yield. Due to
the extraction of heat through the mould walls, the
mush near the walls is likely to reach a low temperature
almost immediately, forming a rigid skin. Shear defects
therefore occur at the edge of the skin where the solid
fraction is less than the maximum packing fraction and
this is the reason for the formation of the bands. The
dierent appearances of the band are caused by the
amount of deformation and the solid fraction when
deformation occurs. Deformation at low solid fractions
results in a highly segregated band, as liquid is concen-

Fig. 13. Representation of the change in shear strength with solid


fraction. The two plots show the dierence in behaviour between ne
globular grains and coarse dendritic grains. The strength increases
rapidly once eutectic soidication begins.

trated to lubricate the ow of the central semisolid


mush. This behaviour is similar to `plug ow' of slurries
where the solid material moves towards the centre of the
ow channel and liquid towards the edge. It is often
observed that the larger coarse dendrites coming from
the shot sleeve segregate towards the centre of a thin
section [8]. With increased solid fraction and deformation there is less liquid available and the amount of
porosity in the band increases, until at the extreme it
appears completely torn [8].
Most of the measurements of the mechanical behaviour of partially solidied material have been undertaken on aluminium alloys [7,39,40]. However,
observation of banded defects is far more common in
magnesium alloys. Also, it is observed that magnesium
alloys can ll large distances along thin sections, much
larger and thinner than that achievable with aluminium
casting alloys. This can be explained by considering the
dierences in the main solidication parameters between, for example, the common Al7 wt% Si and AZ91
casting alloys. There is a signicant dierence in the
freezing range, approximately ve times, between the
two alloys; about 30C for Al7 wt% Si and 160C for
AZ91. The volume fraction of eutectic is about 50% for
Al7 wt% Si and about 25% for AZ91. Fluidity measurements show that AZ91 is much better than Al7
wt% Si and the thermal conductivity of magnesium is
also much lower than that for aluminium. By consideration of the dierences in both the amount of solid
fraction formed and the freezing range before eutectic
solidication begins, it can be realised that AZ91 can
ow signicant distances before the microstructure locks
up when eutectic solidication occurs in the segregated
band. Also, the time to coherency and maximum
packing is greater. Therefore, once a segregated band of
liquid forms, much more ow is possible before this
band solidies. Eventually the temperature drops,
leading to the formation of pores, and then tearing, due
to a lack of liquid to adjust for shrinkage.

A.K. Dahle et al. / Journal of Light Metals 1 (2001) 6172

Whether or not these banded defects form depends on


the volume of ow and the thermal state of the mush
when ow stops. By controlling the casting parameters
and die design it is possible to minimise the degree of
damage caused by these defects [8]. How all these factors
interact is dicult to determine, as there are very limited
data available on the mechanical and physical properties
of magnesium alloys. This is an area of research that
requires signicant attention.

71

mechanical behaviour and solidication properties of


MgAl alloys in order to develop a capability to predict,
and then avoid, the formation of these defects.

Acknowledgements
CAST was established under the Australian Government's Cooperative Research Centres Scheme.

6. Summary
The solidication of magnesiumaluminium alloys
begins with the nucleation of a-Mg dendrites exhibiting
sixfold symmetry. The grain size is set by a combination
of the cooling conditions, alloy composition and the
type of nucleant particles present. The eect of composition on grain size can be estimated by the value of the
growth restriction factor for the alloy. However, alloys
containing aluminium do not exhibit the ne grain sizes
achievable in other magnesium alloys. Also, the grain
size is aected by the purity of the base magnesium used
to make the alloys. It is speculated that aluminium in
combination with manganese and the impurity elements
aects the potency of the nucleants presents in the melt.
The types of nucleant particles naturally occurring in the
melt have not been identied and work has shown that
there are many particles that can be deliberately added
which facilitate nucleation to some extent. However, a
reliable, easy to use, commercial grain rener still needs
to be developed.
After solidication of the a-Mg dendrites, eutectic
solidication occurs as divorced or partially divorced
b-Mg17 Al12 in the interdendritic and grain boundary
regions, surrounded by eutectic a-Mg which is enriched
in aluminium. The degree of divorced growth is aected
by the zinc content and cooling rate, with an increase in
either leading to a more divorced microstructure. The
sensitivity of the eutectic microstructure to ternary elements suggests that further research could be undertaken to manipulate the distribution and morphology of
the b-phase to gain an improvement in as-cast properties. At slow cooling rates, the a-Mg that forms late in
the solidication process decomposes to an a and b
lamellar structure by discontinuous precipitation. Some
continuous precipitation may also occur. Heat treatment is able to completely dissolve the b-phase.
Due to the large freezing range and low eutectic volume fraction of MgAl casting alloys, they can be cast
into large thin sections. However, this advantage also
increases their susceptibility to banded defects of segregated eutectic, porosity or tears. By controlling the
casting parameters the degree of damage caused by these
defects can be minimised. However, much more research
needs to be carried out to gain data on the mushy zone

References
[1] A.K. Dahle, D.H. StJohn, G.L. Dunlop, Developments and
challenges in the utilisation of magnesium alloys, Mater. Forum
24 (2000) 167182.
[2] D.J. Sakkinen, SAE Trans. 103 (1994) 558569.
[3] ASM Handbook Committee, Alloy Phase Diagrams ASM
International, Ohio, USA, 1992.
[4] M.D. Nave, A.K. Dahle, D.H. StJohn, Magnesium technology
2000, in: H.I. Kaplan, J.N. Hryn, B.B. Clow (Eds.), The Minerals,
Metals and Materials Society (TMS), Warrendale, PA, USA,
2000, pp. 233242.
[5] M.D. Nave, A.K. Dahle, D.H. StJohn, Magnesium technology
2000, in: H.I. Kaplan, J.N. Hryn, B.B. Clow (Eds.), The Minerals,
Metals and Materials Society (TMS), Warrendale, PA, USA,
2000, pp. 243250.
[6] S. Sannes, H. Westengen, in: B.L. Mordike, K.U. Kainer (Eds.),
Magnesium alloys and their applications, Werksto-Informationsgesellschaft mbH, Frankfurt, Germany, 1998, pp. 223228.
[7] A.K. Dahle, D.H. StJohn, Acta Mater. 47 (1999) 3141.
[8] A.K. Dahle, D.H. StJohn, Transactions of 20th International Die
Casting Congress & Exposition North American Die Casting
Association (NADCA), November 14, 1999, Cleveland (OH),
Paper #T99062, pp. 203210.
[9] A.L. Bowles, J.R. Griths, C.J. Davidson, Magnesium Technology 2001, The Minerals, Metals and Materials Society (TMS),
Warrendale, PA, USA, 2001, in press.
[10] I.J. Polmear, Light Alloys, Chapman and Hall, London, 1989.
[11] A. Luo, in: G.W. Lorimer (Ed.), Third International Magnesium
Conference, Manchester, UK, 1996, pp. 449464.
[12] C.E. Nelson, AFS Trans. 56 (1948) 123.
[13] R.T. Wood, The Foundryman (1953) 98.
[14] R. Hultgren, D.W. Mitchell, Trans. AIME 161 (1945) 323327.
[15] A. Luo, Scr. Metall. Mater. 31 (1994) 12531258.
[16] D.O. Karlsen, D. Oymo, H. Westengen, P.M.D. Pinfold, S.I.
Stromhaug, Light Metals Processing and Applications Canadian
Institute of Mining, Metallurgy and Petroleum, Quebec City,
Canada, 1993, pp. 397408.
[17] J.E.C. Hutt, D.H. StJohn, L. Hogan, A.K. Dahle, Mater. Sci.
Technol. 15 (5) (1999) 495500.
[18] J.E.C. Hutt, A.K. Dahle, Y.C. Lee, D.H. StJohn, Light Metals
1999, in: C.E. Eckert (Ed.), The Minerals, Metals and Materials
Society (TMS), Warrendale, PA, USA, 1999, pp. 685692.
[19] M. Easton, D. StJohn, Metall. Mater. Trans. A 30A (1999) 1625
1633.
[20] M. Easton, D. StJohn, Acta. Mater. in press.
[21] J. Hunt, Mater. Sci. Eng. 65 (1984) 7583.
[22] Y.C. Lee, A.K. Dahle, D.H. StJohn, Metall. Mater. Trans. A. in
press.
[23] Y.C. Lee, A.K. Dahle, D.H. StJohn, Magnesium Technology
2000, in: H.I. Kaplan, J.N. Hryn, B.B. Clow (Eds.), The Minerals,
Metals and Materials Society, 2000, pp. 211218.

72

A.K. Dahle et al. / Journal of Light Metals 1 (2001) 6172

[24] Y. Tamura, T. Motegi, N. Kono, E. Sato, Mater. Sci. Forum


350351 (2000) 199204.
[25] M.S. Dargusch, G.L. Dunlop, K. Pettersen, in: B.L. Mordike,
K.U. Kainer (Eds.), Magnesium Alloys and Their Applications,
Werksto-Informationsgesellschaft mbH, Wolfsburg, Germany,
1998, pp. 277282.
[26] M. Regev, E. Aghion, A. Rosen, M. Bamberger, Mater. Sci. Eng.
A 252 (1998) 616.
[27] E.M. Gutman, Y. Unigovski, M. Levkovich, Z. Koren,
E. Aghion, M. Dangur, Mater. Sci. Eng. A 234236 (1997)
880883.
[28] S. Guldberg, Ph.D. Thesis, The Norwegian University of Science
and Technology, Trondheium, Norway, 1997.
[29] F.A. Fox, J. Inst. Metals 71 (1945) 415439.
[30] W.A. Baker, J. Inst. Metals 71 (1945) 165204.
[31] E.A.G. Liddard, W.A. Baker, Trans. Amer. Foundrymen's Soc.
53 (1945) 5465.

[32] B.A. Mikucki, J.D. Shearouse, Magnesium Properties and


Applications for Automobiles (1993) 107115.
[33] E.J. Whittenberger, F.N. Rhines, J. Metals (1952) 409420.
[34] E. vrelid, P. Bakke, T.A. Engh, Light Metals (1997) 141154.
[35] P.L. Schaer, Y.C. Lee, A.K. Dahle, Magnesium Technology
2001, in: J. Hryn (Ed.), The Minerals, Metals and Materials
Society, in press.
[36] R.A. Dodd, W.A. Pollard, J.W. Meier, AFS Trans. 65 (1957)
100117.
[37] J. Campbell, Trans. Met. Soc. AIME 245 (1969) 23252334.
[38] W.P. Sequeira, M.T. Murray, G.L. Dunlop, D.H. StJohn, in:
S.K. Das (Ed.), Automotive Alloys 1997, TMS, Warrendale, PA,
1997, pp. 169183.
[39] A.K. Dahle, L. Arnberg, Acta. Mater. 45 (2) (1997) 547559.
[40] T. Sumitomo, D.H. StJohn, T. Steinberg, Mat Sci. Eng. A 289
(2000) 1829.

Potrebbero piacerti anche