Sie sind auf Seite 1von 11

Journal of Materials Processing Technology 210 (2010) 1249–1259

Contents lists available at ScienceDirect

Journal of Materials Processing Technology


journal homepage: www.elsevier.com/locate/jmatprotec

Review

The heat treatment of Al–Si–Cu–Mg casting alloys


Emma Sjölander ∗ , Salem Seifeddine
Materials and Manufacturing – Casting, Department of Mechanical Engineering, School of Engineering, Jönköping University,
Box 1026, SE-551 11 Jönköping, Sweden

a r t i c l e i n f o a b s t r a c t

Article history: The mechanical properties of aluminium–silicon casting alloys containing Cu and Mg are known to be
Received 30 November 2009 improved by heat treatment. Over 60 papers are reviewed here in order to clarify the sequences of
Received in revised form 17 March 2010 microstructural changes which occur during heat treatment, and their influence on the mechanical prop-
Accepted 21 March 2010
erties. It is shown that the changes occurring during solution treatment are relatively well understood,
and that the equilibrium phase diagram can be used to predict the stability of phases at the solution
treatment temperature. The influence of quench rate and natural ageing on subsequent artificial ageing
Keywords:
needs to be studied further, but some conclusions can be drawn. These include: (1) An increase in quench
Cast aluminium alloys
Heat treatment
rate above 4 ◦ C/s gives a small increase in yield strength after ageing, while the concomitant influence
Mechanical properties on elongation is more complicated and depends on the alloy. (2) Natural ageing is shown to have a
large influence on subsequent artificial ageing response of Al–Si–Mg alloys, while there is a significant
lack of knowledge for Cu-containing alloys. Artificial ageing of Al–Si–Mg alloys in the temperature range
170–210 ◦ C gives peak yield strengths of the same level, while Cu-containing alloys show a decrease in
peak yield strength with increasing ageing temperature. The precipitation sequences in Al–Si–Mg and
Al–Si–Cu alloys are relatively well known. In Al–Si–Cu–Mg alloys several precipitation sequences are
possible, which need further investigation.
© 2010 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1250
2. Solution treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1250
2.1. Dissolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1250
2.2. Homogenization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1250
2.3. Examples of dissolution and homogenization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1250
2.3.1. Al–Si–Mg alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1250
2.3.2. Al–Si–Cu alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1251
2.3.3. Al–Si–Cu–Mg alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1251
2.4. Two-stage solution treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1252
2.5. Spheroidization of eutectic Si particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1252
3. Quenching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1253
3.1. Quench sensitivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1253
4. Ageing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1254
4.1. Natural ageing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1254
4.2. Artificial ageing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1254
4.2.1. Precipitation sequence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1254
4.2.2. Hardening mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1254
4.2.3. Al–Si–Mg alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1254
4.2.4. Al–Si–Cu alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1255
4.2.5. Al–Si–Cu–Mg alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1256

∗ Corresponding author.
E-mail address: emma.sjolander@jth.hj.se (E. Sjölander).

0924-0136/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.jmatprotec.2010.03.020
1250 E. Sjölander, S. Seifeddine / Journal of Materials Processing Technology 210 (2010) 1249–1259

4.3. Effect of natural ageing on mechanical properties obtained after artificial ageing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1257
4.3.1. Al–Si–Mg alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1257
4.3.2. Al–Si–Cu alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1257
5. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1258
Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1258
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1258

1. Introduction rate to the solution treatment temperature. The former affects the
fraction, size, morphology and type of Cu-containing phases and
The mechanical properties of cast Al–Si–Cu–Mg alloys depend the latter affects the time available for dissolution of Cu-containing
mainly on the alloy composition and the parameters of the cast- phases during heating.
ing process. In order to further improve the mechanical properties
of cast components these alloys can be heat treated. Various heat 2.1. Dissolution
treatment cycles, e.g. different combinations of temperatures and
times, are used depending on the casting process, the alloy com- Not all phases formed during solidification dissolve during
position and the desired mechanical properties. The Aluminum solution treatment. For example, ␤-Mg2 Si and ␪-Al2 Cu parti-
Association has standardized the definitions and nomenclature for cles are relatively easy to dissolve, while ␲-Al8 Mg3 FeSi6 and
heat treatment. A typical heat treatment applied to sand and grav- Q-Al5 Cu2 Mg8 Si6 particles are hard to dissolve, or transform in
ity die-cast Al–Si alloys is the T6 heat treatment, which involves the solid state (Moustafa et al., 2003). Phases containing Fe are
the following stages: hard to dissolve; and the ␣-Al15 (Fe,Mn)3 Si2 script phase is virtu-
ally unaffected by solution treatment, while the ␤-Al5 FeSi platelets
1. Solution treatment at a relatively high temperature to dissolve fragment and undergo gradual dissolution after a long time at a high
Cu- and Mg-rich particles formed during solidification to achieve temperature (Crowell and Shivkumar, 1995). To realise the full age-
a high and homogeneous concentration of the alloying elements ing potential of the alloy it is important that Mg- and Cu-containing
in solid solution. phases dissolve. Cu and Mg atoms that are bound to phases that do
2. Quenching, usually to room temperature, to obtain a supersatu- not dissolve during solution treatment are not available to increase
rated solid solution of solute atoms and vacancies. the strength by precipitation hardening.
3. Age hardening, to cause precipitation from the supersaturated
solid solution, either at room temperature (natural ageing) or at 2.2. Homogenization
an elevated temperature (artificial ageing).
When atoms detach from the coarse particles formed during
Basic principles of, and microstructural changes occurring dur- solidification they diffuse through the matrix to decrease the con-
ing these three stages, including examples for Al–Si–Mg, Al–Si–Cu centration gradient, forming a homogenous solid solution. The time
and Al–Si–Cu–Mg alloys are given in the following chapters. needed to homogenize the casting is determined by the nature of
the diffusing atoms and the solution treatment temperature (diffu-
2. Solution treatment sion rate) as well as by the diffusing distance which is given by the
coarseness of the microstructure normally measured by secondary
Solution treatment is carried out at a high temperature, close to dendrite arm spacing (SDAS).
the eutectic temperature of the alloy. The purpose of the solution
heat treatment is to: 2.3. Examples of dissolution and homogenization

• dissolve soluble phases containing Cu and Mg formed during The time needed for dissolution and homogenization depends
solidification; on the composition, morphology, size and distribution of the phases
• homogenize the alloying elements; present after solidification, as well as on the solution treatment
• spheroidize the eutectic Si particles. temperature. The phases that can form during solidification for
different alloys are discussed below, as well as typical solution
The rate of these three processes increases as the solu- treatment times and temperatures.
tion treatment temperature increases. The strength that can be
obtained after ageing also increases as the temperature increases, 2.3.1. Al–Si–Mg alloys
because the maximum solubility of solute obtainable in the matrix 2.3.1.1. As-cast condition. Mg-containing phases that can form dur-
increases. The maximum solution treatment temperature that can ing solidification are Mg2 Si and the ␲-Fe phase. The concentrations
be used depends on the Cu and Mg concentrations of the alloy of Mg and Fe in the alloy and the solidification rate determine the
and is limited by incipient melting of phases formed from the last fractions of the ␤-Fe and ␲-Fe phases, as well as the concentra-
solidified melt that is rich in solute elements due to segregation. tion of Mg in solid solution. The ␲-Fe phase has a Chinese script or
Localized melting results in distortion and substantially reduced blocky morphology and is often formed on the ␤-Fe plates. Taylor et
mechanical properties. Cast Al–Si–Mg alloys can be solution treated al. (2000b) report that the Mg level has no significant influence on
at 540–550 ◦ C (Shivkumar et al., 1990b), while alloys containing Cu the volume fraction of the phases formed for an Al–Si–Mg alloy with
must be solution treated at a lower temperature due to the risk of 0.12 wt.% Fe and a SDAS of 40 ␮m. The ␲-Fe phase is the dominant
local melting of Cu-containing phases. According to Samuel (1998) phase, at about 1 vol.%, while the ␤-Fe phase is less than 0.1 vol.%.
Cu-containing phases start to melt at 519 ◦ C in an A319 alloy with The fraction of Mg2 Si phase increases with increasing Mg content
low-Mg concentration, while melting starts at 505 ◦ C in an A319 from close to zero for 0.3 wt.% Mg to 0.2 vol.% for 0.7 wt.% Mg.
alloy with 0.5 wt.% Mg, due to the presence of the Q-Al5 Mg8 Si6 Cu2
phase. The exact temperatures that can be used without localized 2.3.1.2. Solution-treated condition. Dissolution of the Mg2 Si phase
melting depend on the casting solidification rate and the heating is a fast process due to the high solution treatment temperature
E. Sjölander, S. Seifeddine / Journal of Materials Processing Technology 210 (2010) 1249–1259 1251

Fig. 1. (a) Eutectic Al2 Cu and (b) blocky Al2 Cu. With kind permission from Springer Science + Business Media: Journal of Material Science, Effect of alloying elements on the
segregation and dissolution of CuAl2 phase in Al–Si–Cu 319 alloys (Li et al., 2003, Fig. 17).

which can be used and the high diffusion rate of Mg in Al. Rometsch A high solidification rate promotes the formation of the eutectic
et al. (1999) studied the solution treatment of an A356 alloy (SDAS Al2 Cu phase, while Sr modification increases the fraction of the
40 ␮m) and an A357 alloy (SDAS 55 ␮m) at a temperature of 540 ◦ C. blocky Al2 Cu phase (Samuel et al., 1996b).
Dissolution of the Mg2 Si phase in the A356 alloy was completed
after 2–4 min and homogenization was completed after 8–15 min.
2.3.2.2. Solution-treated condition. Samuel et al. (1996a), Li et al.
For the A357 alloy both dissolution and homogenization were com-
(2003) and Han et al. (2008) studied the dissolution of the blocky
plete within 50 min. According to these authors the longer time
Al2 Cu phase and the eutectic Al2 Cu phase during solution treat-
needed for the A357 alloy is due to the higher Mg concentration
ment. The eutectic Al2 Cu dissolves by fragmentation into smaller
and the coarser microstructure. Several other investigations report
segments that spheroidize and finally dissolve by radial diffusion of
similar results, for example Closset et al. (1986), Shivkumar et al.
Cu into the surrounding matrix, see Fig. 2. The blocky Al2 Cu phase
(1990b) and Zhang et al. (2002).
is harder to dissolve than the eutectic Al2 Cu phase, as the former
Dons et al. (2000), Taylor et al. (2000b), Rometsch et al. (2001)
does not fragment but dissolves by spheroidization and diffusion,
and Wang and Davidson (2001) have studied the transformation of
which takes a longer time. The concentration of Al and Cu in the
the ␲-Fe phase during solution treatment. They concluded that the
Al2 Cu particles is constant during the dissolution process and Cu
ability to transform the ␲-Fe phase into ␤-Fe phase and Mg in solid
diffuses from the outer layer of the Al2 Cu particles into the matrix.
solution depends on the Mg concentration of the alloy. If the Mg
Dissolution of Al2 Cu phases takes several hours, due to the low dif-
concentration is low (0.3–0.4 wt.%) the transformation is fast. On
fusion rate of Cu in Al and the low solution treatment temperature
the other hand, if the Mg concentration is high (0.6–0.7 wt.%) there
allowed. Samuel et al. (1996a) for example report that 8 h at 515 ◦ C
will be no transformation and the process may even be reversed. As
or 24 h at 505 ◦ C was needed to dissolve 75% of the Al2 Cu phase in
a result, an increase in Mg level above about 0.5 wt.% is not expected
a 319.2 alloy with SDAS of 22 ␮m.
to give an increase in strength when a T6 heat treatment with a
According to Dons (2001) the ␤-Fe phase transforms into the
solution treatment at 540 ◦ C is used.
Al7 FeCu2 phase during solution treatment when the alloy has a
high Cu concentration. This means that Cu in solid solution is lost
to the Al7 FeCu2 phase. An AlFeSiCu phase has also been reported by
2.3.2. Al–Si–Cu alloys
Cerri et al. (2000) to be present after solution treatment. This phase
2.3.2.1. As-cast condition. The solidification of the Al2 Cu phase has
has the shape of long thin needles and might be an intermediate
been extensively studied, by for example Samuel et al. (1996b),
step in the ␤-Fe to Al7 FeCu2 phase transition.
Djurdjevic et al. (1999) and Li et al. (2003). In the as-cast condi-
tion the Al2 Cu phase can be present in different shapes, as compact
block-like Al2 Cu phase, as eutectic Al2 Cu phase, or as a mixture of 2.3.3. Al–Si–Cu–Mg alloys
both types, see Fig. 1. The Al2 Cu phase nucleates on ␤-Fe plates or 2.3.3.1. As-cast condition. The influence of Mg additions to a 319
on coarse eutectic Si particles during the last stage of solidification. alloy was studied by Samuel et al. (1997). The addition of Mg leads

Fig. 2. Dissolution process of (a) eutectic Al2 Cu and (b) blocky Al2 Cu particles (Han et al., 2008). Reprinted with permission of the American Foundry Society, Schaumburg,
IL, USA (www.afsinc.org).
1252 E. Sjölander, S. Seifeddine / Journal of Materials Processing Technology 210 (2010) 1249–1259

Fig. 3. Eutectic silicon in as-cast Al–11Si–0.6Mg alloy (a) unmodified and (b) Sr modified. With kind permission from Springer Science + Business Media: Metallurgical and
Materials Transaction A, The effect of solution heat treatment and quenching rates on mechanical properties and microstructures in AlSiMg foundry alloys (Pedersen and
Arnberg, 2001, Fig. 7).

to segregation of Cu away from the Si eutectic, resulting in the for- will be a lower concentration of alloying elements and vacancies in
mation of the blocky Al2 Cu phase and the Q-Al5 Mg8 Si6 Cu2 phase solid solution and there will be less spheroidization of Si particles.
growing from the blocky Al2 Cu phase during the last stage of solid- Samuel (1998) for example reports that an addition of Mg to an
ification. The coarseness of the Q phase increases with increasing Al–Si–Cu alloy did not provide an increase in strength at a solution
Mg level. The ␤-Fe phase starts to transform to the ␲-Fe phase as treatment temperature of 480 ◦ C, but when the temperature was
0.35 wt.% Mg is added and the size and volume fraction of the ␲-Fe increased to 500 ◦ C an increase in strength was achieved.
phase increases with increasing Mg. Mg2 Si precipitates as small Sokolowski et al. (1995) investigated the possibility to use a
dots on the Si particles for Mg levels below 0.5 wt.% and in the two-stage solution treatment. The alloy is first solution treated at a
form of Chinese script for Mg levels above 1 wt.%. Alloys with a low temperature, 495 ◦ C for 8 h, to dissolve Cu-containing particles,
low Cu level of 0.5 wt.% behave as an Al–Si–Mg alloy and the Mg2 Si and then the temperature is increased to 520 ◦ C for 2 h to obtain a
phase and the ␲-Fe phase are formed with traces of the Q and homogenous concentration of alloying elements. This procedure
Al2 Cu phase (Dons, 2001). If the Cu and Mg levels are increased to resulted in an increase in both strength and ductility. If too high a
1.4 wt.% the Mg2 Si and ␲-Fe phase are still the main phases formed, temperature is used localized melting occurs and the mechanical
although the amount of Al2 Cu phase and Q phases increase (Lasa properties decrease drastically.
and Rodriguez-Ibabe, 2004).

2.5. Spheroidization of eutectic Si particles


2.3.3.2. Solution treatment condition. The alloy composition and
the solution treatment temperature determine which phases will The third aim of a solution treatment is to spheroidize eutectic
dissolve during solution treatment. The Q phase can for example silicon particles. Apelian et al. (1989) have summarised much of the
be stable, grow or dissolve during solution treatment depending work done on the influence of the solution treatment on the mor-
on the alloy composition and solution treatment temperature. Lasa phology of the eutectic Si particles up to 1989. The morphology of
and Rodriguez-Ibabe (2002) and Han et al. (2008) measured the the eutectic Si has a great impact on the mechanical properties of
area fraction of the Q phase in the as-cast and solution-treated the alloy. Si is present as large brittle flakes in unmodified alloys,
condition of alloys with high Cu (3.5–4.4 wt.%) and various Mg con- see Fig. 3a, which act as crack initiators and have a negative effect
centrations. The area fraction of Q phase did not change during on the alloy ductility. The eutectic Si morphology can be altered,
solution treatment, indicating that the phase is stable or dissolves either by exposing the casting to a high temperature for long peri-
very slowly. Lasa and Rodriguez-Ibabe (2004) on the other hand ods, or by chemical modification by addition of for example Sr to
report an increase in Q phase during solution treatment by substi- the melt, or by a combination of both treatments. During solution
tution of the dissolving Mg2 Si phase by the Q phase in an alloy with treatment the eutectic Si particles first fragment and spheroidize
low Cu (1.4 wt.%) and high Mg (1.3 wt.%) concentrations. In these and then coarsen. When Sr is added the morphology of the eutectic
examples solution treatment temperatures of about 490 ◦ C were Si changes to fibrous, see Fig. 3b. The fibrous morphology is much
used. Lasa and Rodriguez-Ibabe (2004) and Chaudhury and Apelian easier to fragment and spheroidize during solution treatment and
(2006) report that the Q phase starts to dissolve when the tem- the solution treatment time can be shortened.
perature is increased to 530 ◦ C. Lasa and Rodriguez-Ibabe (2004) The time needed for spheroidization depends strongly on the
conclude that dissolution and formation of phases during solution solution treatment temperature and on the morphology and size
treatment is well predicted by the equilibrium phase diagram. of the eutectic Si particles in the as-cast condition. The times pre-
The time needed for dissolution and homogenization is reported sented below give mechanical properties that are close to the
to be shorter for Al–Si–Cu–Mg alloys than for corresponding Mg- maximum values achieved after long solution treatment times
free alloys. According to Han et al. (2008), 8 h at 490 ◦ C is needed and are not only based on visual inspection of the Si particles.
to obtain a high and uniform Cu concentration in the matrix of an Shivkumar et al. (1990b) report that 3–6 h at 540 ◦ C is the optimal
Al–7Si–3.5Cu alloy with SDAS 40–50 ␮m, while only 4 h is needed time for a Sr-modified sand-cast A356 alloy. The time can be short-
if Mg is added to the alloy. ened further when the microstructure is finer. According to Zhang
et al. (2002) 30 min at 540 ◦ C is needed for a low-pressure die-cast
2.4. Two-stage solution treatment Sr-modified Al–7Si–0.3Mg alloy with SDAS of 25 ␮m. Ogris et al.
(2002) report that the time can be decreased to 3 min at 540 ◦ C for a
Alloys containing Cu, or both Cu and Mg cannot be solution thixocast Sr-modified A356 alloy. When the solution temperature
treated at as high temperature as alloys containing only Mg due is lower, as for Al–Si–Cu alloys longer times are needed. Crowell
to the risk of melting the Cu-containing phases. A lower temper- and Shivkumar (1995) report that 8–16 h at 495 ◦ C are needed for
ature will not give an optimal solution treatment because there a Sr-modified 319.1 alloy.
E. Sjölander, S. Seifeddine / Journal of Materials Processing Technology 210 (2010) 1249–1259 1253

Fig. 4. (a) Yield strength and (b) elongation to fracture for T6 heat treated Al–Si alloys subjected to different quench rates. (A) Emadi et al. (2003), (B) Seifeddine et al. (2007),
(C) Zhang and Zheng (1996), (D) Rometsch and Schaffer (2000) and (E) Pedersen and Arnberg (2001).

3. Quenching Mg concentration and much lower, about one-fourth of the quench


sensitivity for quench rates below 1 ◦ C/s. The decrease in strength
The objective of quenching is to suppress precipitation upon with decreasing quench rate is according to Zhang and Zheng
cooling of the casting from the high solution treatment temperature (1996) due to loss of excess Si as the amount of excess Si deter-
to room temperature. If the quench rate is sufficiently high solute mines the composition of the ␤ phase that forms during ageing.
is retained in solid solution and a high number of vacancies are also Less excess Si after quenching thereby results in a lower volume
retained. On the other hand if the cooling is too slow, particles pre- fraction of ␤ precipitates, giving a lower strength.
cipitate heterogeneously at grain boundaries or dislocations, which Few investigations have been reported on Cu-containing alloys.
results in a reduction in supersaturation of solute and concomi- The quench sensitivity for 319 alloys seems to be of the same order
tantly a lower maximum yield strength after ageing. The drawback as for Al–Si–Mg alloys, according to data reported by Garcia-Celis et
with quick cooling is that thermal stresses are induced in the cast- al. (1998). Newkirk et al. (2002) report a higher quench sensitivity of
ing. Water is often used as quenching medium. When a slower an Al–6Si–3.5Cu–0.3Mg alloy compared to an Al–7Si–0.35Mg alloy.
quench rate is needed other quenching media such as oil, salt baths Fig. 4b shows the influence of the quench rate on the elongation
and organic solutions can be used. to fracture. Two different behaviours can be observed. For alloys
Precipitation kinetics depend on the degree of supersaturation with high Mg concentration (0.6 wt.%) the elongation increases
and on the diffusion rate, which vary with temperature in opposite when the quench rate decreases. For alloys with low-Mg concentra-
ways. At a high temperature the diffusion rate is high and the super- tion (0.2–0.4 wt.%) the behaviour is more complex. The elongation
saturation is low, while at a low temperature the diffusion rate decreases with decreasing quench rate until it reaches a minimum
is low and the supersaturation is high. The maximum nucleation at 4–40 ◦ C/s and then increases on further decrease in quench rate.
and growth rates occur over a critical temperature range which is Pedersen and Arnberg (2001) suggest that the ductility is related
between 450 and 200 ◦ C for most Al-alloys. The time spent in this to the amount of excess Si in the matrix after quenching. The
temperature region during quenching should therefore be as short increase in ductility for alloys with higher Mg concentrations is
as possible to avoid precipitation. due to the formation of a lower fraction of coherent ␤ during
ageing after a slow quench. Lower fractions of coherent ␤ have
3.1. Quench sensitivity also been observed in low-Mg alloys, but brittle Si precipitates
are also formed after a slow quench, due to the higher amount of
The quench sensitivity is higher for cast alloys than for wrought excess Si, resulting in a reduction in ductility. Zhang and Zheng
alloys. According to Tiryakioglu and Shuey (2007) the higher (1996) observed a decrease in ductility when quenching at 0.5 ◦ C/s
quench sensitivity of the former is due to the presence of eutec- compared to 110 ◦ C/s and have a slightly different explanation. Si
tic Si particles in cast alloys, which influence the quench sensitivity precipitates were formed in the ␣-Al matrix after ageing when a
in several ways. Firstly, the amount of excess (solid solution) Si in quench rate of 110 ◦ C/s was used. At the extremely slow quench
the matrix is reduced with reduced quench rates due to diffusion rate, 0.5 ◦ C/s, coarse ␤ rods surrounded with precipitate-free zones
of Si atoms to eutectic Si particles. Secondly, the ␤-Mg2 Si phase were formed, while no Si particles were observed in the ␣-Al
nucleates on the eutectic Si particles, reducing the concentration matrix. The coarse ␤ rods are brittle and were thought to be the
of Mg in solid solution. Finally, the presence of eutectic Si particles reason for the reduced ductility of the slow quench.
gives a high dislocation density due to the difference in thermal Zhang and Zheng (1996) and Jeyakumar et al. (2009) report that
expansion between Si and Al, which provides nucleation sites for the quench rate has a strong influence on the ageing curve, with a
precipitates. much reduced increase in hardness for air quenched samples than
The influence of quench rate on yield strength and elongation for water quenched ones. Zhang and Zheng (1996) report that the
are summarised in Fig. 4. All samples were aged 6–8 h at 170 ◦ C after time to peak hardness increases for extremely slow quench rates
quenching, except the samples of Pedersen and Arnberg (2001) (0.5 ◦ C/s), while for faster quench rates, above 20 ◦ C/s, no shift is
which were aged 4 h at 150 ◦ C. The slope of the curves in Fig. 4a seen in the time to the peak. Jeyakumar et al. (2009) studied the
is a measure of the quench sensitivity of the alloy. At low quench influence of natural ageing prior to artificial ageing, on the quench
rates (below 1 ◦ C/s), the yield strength increases linearly with the sensitivity. The difference between the ageing curves for the two
logarithm of the quench rate and does not seem to depend on the quench rates decreased when natural ageing was used, i.e. the
composition of the alloy. At quench rates between 1 and 4 ◦ C/s quench sensitivity is lower when natural ageing is applied prior
the quench sensitivity is alloy dependent, and quench sensitiv- to artificial ageing. Quench sensitivity is also affected by the ageing
ity increases with increasing Mg concentration. At higher quench condition, as reported by Zhang and Zheng (1996) and Newkirk et
rates (above 4 ◦ C/s), the quench sensitivity is independent of the al. (2002).
1254 E. Sjölander, S. Seifeddine / Journal of Materials Processing Technology 210 (2010) 1249–1259

Fig. 5. Natural ageing curves for (a) SSM HPDC A356 alloy (Möller et al., 2007) and (b) an Al–9Si–3.5Cu and an Al–9Si–3.5Cu–0.5Mg alloy (Reif et al., 1997b).

4. Ageing The process is driven by a reduction in surface energy, so that the


larger precipitates coarsen as the smaller ones dissolve. As the pre-
Ageing takes place at room temperature (natural ageing) or at an cipitates grow the coherency strain increases until interfacial bond
elevated temperature in the range of 150–210 ◦ C (artificial ageing). strength is exceeded and the precipitates become non-coherent.
The objective of ageing is to obtain a uniform distribution of small Last to form in the precipitation sequence is the non-coherent equi-
precipitates, which gives a high strength. librium phase. The precipitation procedure does not necessarily
follow the above sequence, it can also start at an intermediate stage,
4.1. Natural ageing depending on the thermal history of the material (natural ageing,
artificial ageing temperature, heating rate, etc.).
The high level of supersaturation and the high vacancy concen-
tration after quenching cause rapid formation of Guinier-Preston 4.2.2. Hardening mechanisms
(GP) zones, which are clusters that contain a high fraction of solute The strength of an alloy derives from the ability of precipitates
atoms. These clusters are very small and finely dispersed in the to stop mobile dislocations. The strength is determined by the size
matrix, because diffusion is limited at room temperature. The clus- and distribution of the precipitates and by the coherency of the pre-
ters are coherent with the matrix, but elastic stresses are induced cipitates with the matrix. The interaction with the dislocations can
around the clusters due to the difference in size between the solute be described by the Friedel effect and by the Orowan mechanism.
and the solvent atoms. Coherent precipitates are formed because Small and not too hard precipitates are normally sheared by moving
they have a small interface energy which gives a smaller critical dislocations (Friedel effect), see Fig. 6a. When the precipitates are
radius that makes them precipitate relatively easy. The GP zones larger and harder the moving dislocations bypass the precipitates
and the stress field around them hinder dislocation motion, result- by bowing (Orowan mechanism), see Fig. 6b. The strength of the
ing in an increase in strength. precipitates increases with their size as long as they are sheared by
Alloys containing Mg harden rapidly at room temperature, see dislocations. Further increase of precipitate size makes the shearing
Fig. 5a. An increase in hardness is seen after about 1 h and a hard- processes rather difficult; thus, it is more favourable for the dislo-
ness plateau is reached after about 100 h, after which no further cations to pass the precipitates via the Orowan mechanism, leading
increase in hardness occurs (Shivkumar et al., 1990a; Möller et al., to a decrease in strength with further increase in precipitate size,
2007). The hardness that can be reached depends on the compo- see Fig. 6c. The highest strength is obtained when there is an equal
sition of the alloy; a higher Mg concentration (<0.7 wt.%) gives a probability for the dislocations to pass the precipitates by shearing
higher hardness. Al–Si–Cu alloys harden slowly at room temper- and by bowing.
ature. If a small concentration of Mg is added the alloy responds
quicker to natural ageing (Reif et al., 1997b), see Fig. 5b.
4.2.3. Al–Si–Mg alloys
The precipitation sequence for Al–Si–Mg alloys starts with the
4.2. Artificial ageing formation of spherical GP zones consisting of an enrichment of Mg
and Si atoms. The zones elongate and develop into a needle shaped
Artificial ageing involves precipitation at an elevated temper- coherent ␤ phase. The needles grow to become semi-coherent
ature, normally in the range 150–210 ◦ C. At these temperatures rods (␤ phase) and finally non-coherent platelets (stable ␤ phase)
atoms can move over larger distances and the precipitates formed (Shivkumar et al., 1990a). Zhang and StJohn (1995) observed a large
during artificial ageing are normally much larger in size than GP number of fine ␤ phases which had a diameter of 2–5 nm and
zones. a length of 10–20 nm in the peak-aged condition. The length of
the needles increased as ageing proceeded. The composition of the
4.2.1. Precipitation sequence metastable precipitates was earlier thought to be close to that of the
The precipitation sequence starts with the formation of GP equilibrium precipitates, but investigations using atom probe field
zones. Next to form are metastable precipitates that are either ion microscopy revealed that the composition may vary. Edwards et
coherent or semi-coherent with the matrix. The metastable precip- al. (1998) for example report that the ␤ precipitate may have Mg:Si
itates may nucleate on the GP zones if they have reached a critical ratios of 1.2 rather than the equilibrium ratio 2. Maruyama et al.
size, or homogeneously in the matrix, or heterogeneously on dislo- (1997) report that the Mg:Si ratio increases through the sequence
cations or other lattice defects. The metastable precipitates grow on GP zones, ␤ , ␤ , ␤ and that the composition of the phases may be
further ageing by diffusion of atoms from the supersaturated solid affected by the excess Si concentration after quenching.
solution to the precipitates. As the supersaturation decreases, the Si is reported by for example Kaczorowski et al. (1979) and
precipitates continue to grow in accordance with Ostwald ripening. Zhang and StJohn (1995) to precipitate in the ␣-Al matrix after
E. Sjölander, S. Seifeddine / Journal of Materials Processing Technology 210 (2010) 1249–1259 1255

Fig. 6. Dislocations pass precipitates either by (a) shearing or (b) bypassing (Gerold, 1979). (c) Relationship between precipitate radius and strength of the particles to resist
shearing or bypassing by dislocations (Polmear, 2006).

slight overageing if Si is present in excess after quenching. The shortened. The time to peak hardness is about 10 h at 170 ◦ C, while
diameter of the Si precipitates formed is about 40 nm (Zhang and it is only 20 min at 210 ◦ C. If temperature is increased above 210 ◦ C
StJohn, 1995). A high concentration of excess Si after quenching is a decrease in strength is seen, see Fig. 8a. According to Eskin (2003)
thought to give a ␤ phase with a low Mg:Si ratio (Zhang, 1996). A the ␤ phase is substituted by the ␤ phase at temperatures above
high fraction of the available Si is then consumed by formation of 200 ◦ C, which gives a lower contribution to strength. The decrease
the ␤ phase and a low concentration of Si remains in solid solu- in peak strength for temperatures above 210 ◦ C could thereby be
tion in the matrix. This concentration is too low for Si precipitates to due to a change in hardening precipitates.
form on initial ageing. The composition of the metastable ␤ precip-
itates changes as ageing proceeds and Si is released into the matrix
and Si precipitates are formed later (Eskin, 2003). 4.2.4. Al–Si–Cu alloys
The peak yield strength increases as the Mg concentration in the GP zones consisting of localized concentrations of Cu atoms,
alloy is increased (Taylor et al., 2000a), see Fig. 7. The increase in having diameters of 3–5 nm, form at room temperature. When
yield strength is linear up to a Mg concentration of around 0.5 wt.%, the temperature is increased above 100 ◦ C these GP zones dis-
while the increase is greatly reduced for higher Mg concentra- solve (Hatch, 1984). In their place ␪ phases (sometimes referred
tions. The ␲-Fe phase is stable at higher Mg concentrations, as to as GP2) are formed. A reduction in strength is seen as the GP
earlier discussed, and cannot be completely dissolved during solu- zones dissolve before the ␪ phase has had time to form. After pro-
tion treatment (Taylor et al., 2000b; Dons, 2001). This means that longed ageing the ␪ phase transforms into the metastable ␪ , which
some of the Mg will be bound to the ␲-Fe phase and cannot be used is partially coherent with the matrix. Finally the stable incoher-
for precipitation hardening, with a lower strength as result. ent equilibrium phase ␪ (Al2 Cu) is formed. Maximum strength is
When artificial ageing temperatures in the range 170–210 ◦ C are obtained when the highest fraction of ␪ phase is present (Hatch,
used for Al–Si–Mg alloys, comparable strength levels are achieved 1984).
(Rometsch and Schaffer, 2002; Alexopoulos and Pantelakis, 2004), A high concentration of dislocations is formed in Al–Si–Cu alloys
see Fig. 8. If a high temperature is used the time for ageing can be during quenching from the solution treatment temperature, due to
the differences in thermal expansion of the ␣-Al matrix and the Si
particles. GP zones form on the dislocations because less activa-
tion energy is needed to form heterogeneously. The GP zones that
form on the dislocations immediately dissolve and change into the
␪ phase (Kang et al., 1999), see Fig. 9a. Cu that should have been
precipitated in fine and uniformly distributed precipitates during
ageing is lost to coarse ␪ phases nucleated on dislocations, result-
ing in lower strength. Reif et al. (1997a) and Wang et al. (2004) only
observed the ␪ phase precipitated on dislocations in the peak-aged
condition, while Kang et al. (1999) observed both ␪ and ␪ pre-
cipitates in the underaged condition. Plate-like ␪ particles with a
length of 100 nm were present on dislocations and fine, uniformly
distributed ␪ precipitates were present in the matrix in regions
without dislocations, see Fig. 9a.
The time to obtain peak hardness is longer for Al–Si–Cu alloys
than for Al–Si–Mg and Al–Si–Cu–Mg alloys. Kang et al. (1999) report
a time to peak hardness of about 120 h at 160 ◦ C and 15 h at 200 ◦ C.
There is a scatter in times to peak hardness reported in the liter-
ature, from 30 h up to 120 h or more at 160 ◦ C (Reif et al., 1997a;
Fig. 7. Yield strength versus nominal Mg content for a peak-aged Al–7Si–Mg alloy Kang et al., 1999; Ouellet and Samuel, 1999; Wang et al., 2004). One
(Taylor et al., 2000a). reason for the scatter is that Mg, even at low concentrations, has a
1256 E. Sjölander, S. Seifeddine / Journal of Materials Processing Technology 210 (2010) 1249–1259

Fig. 8. Artificial ageing curves for two A357 alloys; (a) Rometsch and Schaffer (2002) and (b) Alexopoulos and Pantelakis (2004).

large influence on the age hardening response. The peak hardness were performed by Li et al. (2006) and Wang et al. (2007), but the
decreases when the ageing temperature is increased, see Fig. 9b, results reported are different. Both added Cu in different concen-
i.e. Al–Si–Cu alloys are more sensitive to ageing temperature than trations (0–4 wt.%) to Al–Si–Mg alloys with around 0.5 wt.% Mg. The
Al–Si–Mg alloys. samples were heat treated to the T6 condition and hardness was
measured. The ageing curves are shown in Fig. 10. According to Li
et al. (2006) the time to peak hardness is the same for Al–Si–Mg
4.2.5. Al–Si–Cu–Mg alloys
and Al–Si–Cu–Mg alloys, indicating that the same kinds of pre-
The strength obtained for Al–Si–Cu–Mg alloys after heat treat-
cipitates are present in both alloys. However, according to Wang
ment is much higher than for the ternary alloys. Ouellet and Samuel
et al. (2007) the time to peak hardness is longer for Al–Si–Cu–Mg
(1999) report that when Mg is added to an Al–Si–Cu alloy the yield
alloys than for Al–Si–Mg alloys, indicating a change in the precipi-
strength increases from 337 to 415 MPa for peak ageing at 150 ◦ C,
tation sequence as Cu is added. This was confirmed by transmission
but the elongations decrease to less than 1%. The ability to age
electron microscopy (TEM) studies on peak-aged material. Li et
harden, that can be measured as the difference in hardness in the as
al. (2006) report that the number density of the ␤ precipitates
quenched condition and in the peak-aged condition, also increases
increased when Cu was added and that plate shaped ␪ phases were
from 170 to 202 MPa as 0.4 wt.% Mg was added. As for Al–Si–Cu
formed in the alloy with 3 wt.% Cu. A few lath-shaped precipitates
alloys the peak strength is reduced for Al–Si–Cu–Mg alloys when
that are precursors to the Q phase were also observed in the alloys
the ageing temperature is increased (Ouellet and Samuel, 1999;
containing Cu. The fraction of the lath-shaped phase increased with
Geier et al., 2006).
ageing time. Wang et al. (2007) on the other hand report that the
Many different precipitates in different combinations have been
hardening precipitate was dot shaped Q phases in the alloy with
observed in the peak-aged condition of Al–Si–Cu–Mg alloys. The
1 wt.% Cu and that ␪ precipitates were formed in addition to the
kind of precipitates that form depends on the alloy composition, the
Q phase for the alloy with 3 wt.% Cu. Wang et al. (2007) found that
thermal history of the alloy and the artificial ageing applied. Exam-
the ability to age harden diminished as 1 wt.% Cu was added to the
ples of precipitates that can be present in the peak-aged condition
Al–Si–Mg alloy due to the formation of the Q phase that has a lower
are: ␤ (Mg2 Si), ␪ (Al2 Cu) and Q (Al5 Mg8 Si6 Cu2 ). It is apparent that
hardenability than the ␤ phase. From these two investigations it is
there are two possible combinations of precipitates. Precipitation
clear that both ␤ and Q phases can form during artificial ageing.
of ␤ and ␪ , or alternatively Q and ␪ , where the ␪ phase forms
The time to peak hardness is shorter and the age-hardenability is
if the Cu concentration is high enough. Two similar investigations

Fig. 9. (a) Precipitation distribution in the matrix for Al–Si–Cu alloys (Kang et al., 1999). (b) Artificial ageing curves for an Al–7Si–3Cu alloy (Kang et al., 1999). Reprinted
with permission of the American Foundry Society, Schaumburg, IL, USA (www.afsinc.org).
E. Sjölander, S. Seifeddine / Journal of Materials Processing Technology 210 (2010) 1249–1259 1257

Fig. 10. Artificial ageing curves for (a) Al–7Si–0.45Mg–xCu at 175 ◦ C (Li et al., 2006) and (b) Al–8Si–0.4Mg–xCu at 160 ◦ C (Wang et al., 2007).

higher for the ␤ phase. Both investigations agree that the ␪ phase a short ageing directly after quenching (Pashley et al., 1966). The
is formed when the Cu concentration is higher than 1 wt.%. natural ageing response can also be decreased by addition of trace
Mishra et al. (2001) performed a study on an Al–339 alloy that elements (Cd, In, Sn and Cu) to an Al–Si–Mg alloy through binding
shows the importance of the thermal history of the alloy on the pre- of vacancies to the trace elements (Ghate et al., 1984).
cipitates formed during artificial ageing. The S phase was observed
in the T5 condition, whereas the ␤ phase formed in the T6 con-
4.3.1. Al–Si–Mg alloys
dition. Similar investigations have been done for wrought alloys,
Several investigations have been made on the effect of natural
and show that natural ageing and artificial ageing have a impor-
ageing of Al–Si–Mg alloys on mechanical properties after artificial
tant influence on the phases formed. The fraction of the Q phase
ageing (Ghate et al., 1984; Shivkumar et al., 1990a; Murali et al.,
formed relative to the ␤ phase is reported to increase with natural
1997; Emadi et al., 2003; Hernández-Paz et al., 2004; Möller et al.,
ageing, ageing time and ageing temperature (Eskin, 2003; Wang et
2007). Möller et al. (2007) report that the peak hardness is shifted to
al., 2003, 2006).
longer ageing times when natural ageing is used compared to direct
ageing after quenching, but the peak strength is not changed, see
Fig. 11. The delayed increase in hardness of the natural aged sam-
4.3. Effect of natural ageing on mechanical properties obtained
ples is thought to be due to dissolution of clusters formed during
after artificial ageing
natural ageing.
Clusters of atoms formed during natural ageing can either
redissolve or continue to grow upon subsequent artificial ageing, 4.3.2. Al–Si–Cu alloys
depending on their size. If the clusters are larger than a critical Few experiments were found in the literature concerning
radius, that depends on the supersaturation, they will be stable the influence of natural ageing on the artificial ageing response
and can act as nucleation points for precipitates (Shivkumar et al., of Al–Si–Cu and Al–Si–Cu–Mg alloys. A positive response was
1990a). The critical radius is larger at a higher temperature as a reported by Geier et al. (2006) for a high pressure die-cast (HPDC)
consequence of the lower supersaturation. If the clusters have not Al–9Si–3Cu–0.3Mg–0.5Zn–0.9Fe alloy in the T5 condition. The time
attained the critical size they will be unstable and redissolve. The to peak strength decreased and the peak strength increased with
solute concentration in the matrix increases and the critical radius natural ageing time, see Fig. 12. These positive effects were seen for
decreases, which will make some of the clusters stable (Jacobs, all artificial ageing temperatures investigated, from 140 to 240 ◦ C.
1999). The larger clusters start to grow and the smaller ones con- The positive response to natural ageing was attributed to the for-
tinue to dissolve. A microstructure with a lower number density mation of both coherent and incoherent precipitates. It should be
of coarser particles is formed compared to the microstructure of noted that the alloy contains Zn, which may affect the artificial age-
directly aged material (Shivkumar et al., 1990a). The detrimental ing response. Reif et al. (1997b) report a slight increase in peak
effect of natural ageing can be nullified by a short high temper- hardness of an Al–9Si–3.5Cu and an Al–9Si9–3.5Cu–0.5Mg alloy
ature treatment prior to artificial ageing (Murali et al., 1997) or upon natural ageing. The small effect of natural ageing on the arti-

Fig. 11. Artificial ageing curves at 180 ◦ C for an A356 alloy with (a) 0 h of natural ageing and (b) 20 h of natural ageing (Möller et al., 2007).
1258 E. Sjölander, S. Seifeddine / Journal of Materials Processing Technology 210 (2010) 1249–1259

Chaudhury, S.K., Apelian, D., 2006. Fluidized bed heat treatment of cast Al–Si–Cu–Mg
alloys. Metall. Mater. Trans. A 37, 2295–2311.
Closset, B., Drew, R.A.L., Gruzleski, J.E., 1986. Eutectic silicon shape control by in situ
measurement of resistivity. AFS Trans. 94, 9–16.
Crowell, N., Shivkumar, S., 1995. Solution treatment effects in cast Al–Si–Cu alloys.
AFS Trans. 107, 721–726.
Djurdjevic, M., Stockwell, T., Sokolowski, J., 1999. The effect of strontium on the
microstructure of the aluminium–silicon and aluminium–copper eutectics in
the 319 aluminium alloy. Int. J. Cast Metal. Res. 12, 67–73.
Dons, A.L., 2001. The Alstruc homogenization model for industrial aluminum alloys.
J. Light Met. 1, 133–149.
Dons, A.L., Pedersen, L., Brusethaug, S., 2000. Modelling the microstructure of heat
treated AlSi foundry alloys. Aluminium 76, 294–297.
Edwards, G.A., Stiller, K., Dunlop, G.L., Couper, M.J., 1998. The precipitation sequence
in Al–Mg–Si alloys. Acta Mater. 46, 3893–3904.
Emadi, D., Whiting, L.V., Sahoo, M., Sokolowski, J.H., Burke, P., Hart, M., 2003. Optimal
heat treatment of A356.2 alloy. In: Crepeau, P.N. (Ed.), Light Metals 2003. TMS,
Fig. 12. Increase in yield strength by combining natural ageing and artificial ageing San Diego, CA, United States, pp. 983–989.
at 200 ◦ C for a HPDC Al–9Si–3Cu–0.3Mg–0.5Zn–0.9Fe alloy (Geier et al., 2006). Eskin, D.G., 2003. Decomposition of supersaturated solid solutions in Al–Cu–Mg–Si
alloys. J. Mater. Sci. 38, 279–290.
Garcia-Celis, A.I., Velasco, E., Valtierra, S., Mojica, J.F., Colas, R., 1998. Cooling effects
on aging in a cast aluminum alloy. In: Das, S.K. (Ed.), TMS Annual Meeting. San
ficial ageing response may be due to the short natural ageing time
Antonio, TX, USA. TMS, pp. 135–143.
of 4 h. Geier, G., Rockenschaub, H., Pabel, T., Hopfinger, M., 2006. Variation of the pre-
cipitation mechanisms for high pressure die casting alloy AlSi9Cu3(Fe)—a
new method of heat treatment for superior mechanical properties.
5. Conclusions Giessereiforschung 58, 32–48.
Gerold, V., 1979. Precipitation hardening. In: Nabarro, F.R.N. (Ed.), Dislocations in
The solution treatment process is relatively well known and Solids. North-Holland, p. 222.
Ghate, G.P., Sreenivasa Murthy, K.S., Raman, K.S., 1984. Effect of trace elements of
the equilibrium phase diagram can be used to predict the stabil- the delayed artificial aging of Al–7% Si–0.3% Mg alloy. Aluminium 60, 19–20.
ity of phases at the solution treatment temperature. When Sr is Han, Y.M., Samuel, A.M., Samuel, F.H., Valtierra, S., Doty, H.W., 2008. Effect of solution
used to modify the eutectic Si, it allows the use of shorter solution heat treatment type on the dissolution of copper phases in Al–Si–Cu–Mg type
alloys. AFS Trans. 116, 79–90.
treatment times, as the spheroidization process is very rapid. Hatch, J.E., 1984. Aluminum: Properties and Physical Metallurgy. ASM, Metals Park,
The quench sensitivity of cast Al–Si alloys is high due to diffu- OH, pp. 136–143.
sion of Si atoms from the matrix to the eutectic Si particles. The Hernández-Paz, J.F., Paray, F., Gruzleski, J.E., Emadi, D., 2004. Natural aging and heat
treatment of A356 aluminum alloy. AFS Trans. 112, 155–164.
strength after ageing increases with increasing quench rate, but
Jacobs, M.H., 1999. TALAT Lecture 1204, Precipitation Hardening
only a small increase in strength is obtained when the quench rate [Online]. EAA, The University of Birmingham, UK, Available:
exceeds about 4 ◦ C/s. The relationship between the elongation after http://www.eaa.net/eaa/education/TALAT/1000/1200.htm.
ageing and the quench rate is more complicated and depends on Jeyakumar, M., Hamed, M., Shankar, S., 2009. Heat treatment of A356.2 aluminum
alloy: effect of quench rate and natural ageing. In: Campbell, J., Crepeau, P.N.,
the alloy composition. Few investigations show the influence of the Tiryakioglu, M. (Eds.), Shape Casting: 3rd International Symposium. San Fran-
quench rate on the artificial ageing curve and there is a lack of data cisco, CA. TMS, pp. 87–95.
for Cu-containing alloys. Kaczorowski, M., Grabski, M.W., Sawicki, J., Murza-Mucha, P., 1979. A study of
precipitation hardening of commercial Al–9 wt% Si alloy. J. Mater. Sci. 14,
Natural ageing is shown to have a large influence on the arti- 2781–2786.
ficial ageing response of Al–Si–Mg alloys, although the peak yield Kang, H.G., Kida, M., Miyahara, H., Ogi, K., 1999. Age-hardening characteristics of
strength and the time to peak is not affected. For Cu-containing Al–Si–Cu-base cast alloys. AFS Trans. 27, 507–515.
Lasa, L., Rodriguez-Ibabe, J.M., 2002. Characterization of the dissolution of the Al2 Cu
alloys the influence is not known. phase in two Al–Si–Cu–Mg casting alloys using calorimetry. Mater. Charact. 48,
Artificial ageing of Al–Si–Mg alloys anywhere in the tempera- 371–378.
ture range 170–210 ◦ C gives the same peak yield strength, while Lasa, L., Rodriguez-Ibabe, J.M., 2004. Evolution of the main intermetallic phases
in Al–Si–Cu–Mg casting alloys during solution treatment. J. Mater. Sci. 39,
Cu-containing alloys show a decrease in peak yield strength with 1343–1355.
increasing ageing temperature. Several precipitation sequences are Li, Y.J., Brusethaug, S., Olsen, A., 2006. Influence of Cu on the mechanical properties
possible in Al–Si–Cu–Mg alloys and it is unknown which parame- and precipitation behavior of AlSi7 Mg0.5 alloy during aging treatment. Scripta
Mater. 54, 99–103.
ters determine the actual sequence which takes place.
Li, Z., Samuel, A.M., Samuel, F.H., Ravindran, C., Valtierra, S., 2003. Effect of alloying
This review shows that it is of vital importance to take the elements on the segregation and dissolution of CuAl2 phase in Al–Si–Cu 319
whole heat treatment process into consideration in order to achieve alloys. J. Mater. Sci. 38, 1203–1218.
the optimal mechanical properties of an alloy. It is not sufficient Maruyama, N., Uemori, R., Hashimoto, N., Saga, M., Kikuchi, M., 1997. Effect of silicon
addition on the composition and structure of fine-scale precipitates in Al–Mg–Si
to consider only the solution treatment and the artificial ageing alloys. Scripta Mater. 36, 89–93.
parameters. Mishra, R.K., Smith, G.W., Baxter, W.J., Sachdev, A.K., Franetovic, V., 2001. The
sequence of precipitation in 339 aluminum castings. J. Mater. Sci. 36, 461–468.
Moustafa, M.A., Samuel, F.H., Doty, H.W., 2003. Effect of solution heat treatment and
Acknowledgement additives on the microstructure of Al–Si (A413.1) automotive alloys. J. Mater.
Sci. 38, 4507–4522.
Murali, S., Arunkumar, Y., Chetty, P.V.J., Raman, K.S., Murthy, K.S.S., 1997. The effect
The European Project NADIA (New Automotive components of preaging on the delayed aging of Al–7Si–0.3Mg. JOM 49, 29–33.
Designed for and manufactured by Intelligent processing of light Möller, H., Govender, G., Stumpf, W.E., 2007. Natural and artificial aging response
Alloys) is gratefully acknowledged for financial support. of semisolid metal processed Al–Si–Mg alloy A356. Int. J. Cast Metal. Res. 20,
340–346.
Newkirk, J.W., Liu, Q., Mohammadi, A., 2002. Optimizing the aging heat treatment of
References cast aluminum alloys. In: Das, S.K., Skillingberg, M.H. (Eds.), TMS Annual Meet-
ing:Automotive Alloys and Aluminum Sheet and Plate Rolling and Finishing
Alexopoulos, N.D., Pantelakis, S.G., 2004. Quality evaluation of A357 cast aluminum Technology Symposia. Seattle, WA, pp. 75–82.
alloy specimens subjected to different artificial aging treatment. Mater. Design Ogris, E., Wahlen, A., Lüchinger, H., Uggowitzer, P.J., 2002. On the silicon spheroidiza-
25, 419–430. tion in Al–Si alloys. J. Light Met. 2, 263–269.
Apelian, D., Shivkumar, S., Sigworth, G., 1989. Fundamental aspects of heat treatment Ouellet, P., Samuel, F.H., 1999. Effect of Mg on the ageing behaviour of Al–Si–Cu 319
of cast Al–Si–Mg alloys. AFS Trans. 137, 727–742. type aluminium casting alloys. J. Mater. Sci. 34, 4671–4697.
Cerri, E., Evangelista, E., Spigarelli, S., Cavaliere, P., DeRiccardis, F., 2000. Effects of Pashley, D.W., Rhodes, J.W., Sendorek, A., 1966. Delayed ageing in
thermal treatments on microstructure and mechanical properties in a thixocast aluminum–magnesium–silicon alloys: effect on structure and mechanical
319 aluminum alloy. Mater. Sci. Eng. A: Struct. 284, 254–260. properties. J. Inst. Met. 94, 41–49.
E. Sjölander, S. Seifeddine / Journal of Materials Processing Technology 210 (2010) 1249–1259 1259

Pedersen, L., Arnberg, L., 2001. The effect of solution heat treatment and quenching Shivkumar, S., Ricci Jr., S., Keller, C., Apelian, D., 1990b. Effect of solution treatment
rates on mechanical properties and microstructures in AlSiMg foundry alloys. parameters on tensile properties of cast aluminum alloys. J. Heat Treating 8,
Metall. Mater. Trans. A 32, 525–532. 63–70.
Polmear, I.J., 2006. Light Alloys; From Traditional Alloys to Nanocrystals. Butter- Sokolowski, J.H., Sun, X.C., Byczynski, G., Northwood, D.O., Penrod, D.E., Thomas, R.,
worth Heinemann, Oxford, pp. 43–69. Esseltine, A., 1995. Removal of copper-phase segregation and the subsequent
Reif, W., Dutkiewicz, J., Ciach, R., Yu, S., Krol, J., 1997a. Effect of ageing on the evolution improvement in mechanical properties of cast 319 aluminum alloys by a two-
of precipitates in AlSiCuMg alloys. Mater. Sci. Eng. A 234-236, 165–168. stage solution heat treatment. J. Mater. Process. Technol. 53, 385–392.
Reif, W., Yu, S., Dutkiewicz, J., Ciach, R., Krol, J., 1997b. Pre-ageing of AlSiCuMg Taylor, J.A., St John, D.H., Barresi, J., Couper, M.J., 2000a. An empirical analysis of
alloys in relation to structure and mechanical properties. Mater. Design 18, trends in mechanical properties of T6 heat treated Al–Si–Mg casting alloys. Int.
253–256. J. Cast Metal. Res. 12, 419–430.
Rometsch, P.A., Arnberg, L., Zhang, D.L., 1999. Modelling dissolution of Mg2 Si and Taylor, J.A., St John, D.H., Barresi, J., Couper, M.J., 2000b. Influence of Mg content on
homogenisation in Al–Si–Mg casting alloys. Int. J. Cast Metal. Res. 12, 1–8. the microstructure and solid solution chemistry of Al–7%Si–Mg casting alloys
Rometsch, P.A., Schaffer, G.B., 2000. Quench modelling of Al–7Si–Mg casting alloys. during solution treatment. Mater. Sci. Forum 331-337, 277–282.
Int. J. Cast Metal. Res. 12, 431–439. Tiryakioglu, M., Shuey, R.T., 2007. Quench sensitivity of an Al–7 pct Si-0.6 pct Mg
Rometsch, P.A., Schaffer, G.B., 2002. An age hardening model for Al–7Si–Mg casting alloy: characterization and modeling. Metall. Mater. Trans. B 38, 575–582.
alloys. Mater. Sci. Eng. A: Struct. 325, 424–434. Wang, G., Bian, X., Liu, X., Zhang, J., 2004. Effect of Mg on age hardening and precip-
Rometsch, P.A., Schaffer, G.B., Taylor, J.A., 2001. Mass balance characterisation of itation behavior of an AlSiCuMg cast alloy. J. Mater. Sci. 39, 2535–2537.
Al–7Si–Mg alloy microstructures as a function of solution treatment time. Int. J. Wang, G., Sun, Q., Feng, L., Hui, L., Jing, C., 2007. Influence of Cu content on ageing
Cast Metal. Res. 14, 59–69. behavior of AlSiMgCu cast alloys. Mater. Design 28, 1001–1005.
Samuel, A.M., Gauthier, J., Samuel, F.H., 1996a. Microstructural aspects of the disso- Wang, Q.G., Davidson, C.J., 2001. Solidification and precipitation behaviour of
lution and melting of Al2 Cu phase in Al–Si alloys during solution heat treatment. Al–Si–Mg casting alloys. J. Mater. Sci. 36, 739–750.
Metall. Mater. Trans. A 27, 1785–1798. Wang, X., Esmaeili, S., Lloyd, D.J., 2006. The sequence of precipitation in the
Samuel, A.M., Ouellet, P., Samuel, F.H., Doty, H.W., 1997. Microstructural interpre- Al–Mg–Si–Cu alloy AA6111. Metall. Mater. Trans. A 37, 2691–2699.
tation of thermal analysis of commercial 319 Al alloy with Mg and Sr additions. Wang, X., Poole, W.J., Esmaeili, S., Lloyd, D.J., Embury, J.D., 2003. Precipitation
AFS Trans. 105, 951–962. strengthening of the aluminum alloy AA6111. Metall. Mater. Trans. A 34,
Samuel, E.H., Samuel, A.M., Doty, H.W., 1996b. Factors controlling the type 2913–2924.
and morphology of Cu-containing phases in 319 Al alloy. AFS Trans. 30, Zhang, D.L., 1996. Precipitation of excess silicon during heat treatment of cast
893–901. Al–7 wt%Si–0.4 wt%Mg alloy. Mater. Sci. Forum 217-222, 771–776.
Samuel, F.H., 1998. Incipient melting of Al5 Mg8 Si6 Cu2 and Al2 Cu intermetallics in Zhang, D.L., StJohn, D.H., 1995. The effect of composition and heat treatment on
unmodified and strontium-modified Al–Si–Cu–Mg (319) alloys during solution the microstructure and mechanical properties of cast Al–7 wt%Si–Mg alloys. In:
heat treatment. J. Mater. Sci. 33, 2283–2297. Casting & Solidification of Light Alloys, Gold Coast, IMMA, pp. 21–25.
Seifeddine, S., Timelli, G., Svensson, I.L., 2007. Influence of quench rate on the Zhang, D.L., Zheng, L., 1996. The quench sensitivity of cast Al–7 Wt Pet Si–0.4 Wt pct
microstructure and mechanical properties of aluminium alloys A356 and A354. Mg alloy. Metall. Mater. Trans. A 27, 3983–3991.
Int. Foundry Res. 59, 2–10. Zhang, D.L., Zheng, L.H., StJohn, D.H., 2002. Effect of a short solution treat-
Shivkumar, S., Keller, C., Apelian, D., 1990a. Aging behavior in cast Al–Si–Mg alloys. ment time on microstructure and mechanical properties of modified
AFS Trans. 98, 905–911. Al–7 wt.%Si–0.3 wt.%Mg alloy. J. Light Met. 2, 27–36.

Das könnte Ihnen auch gefallen