You are on page 1of 7

Materials Science and Engineering A301 (2001) 140 146

www.elsevier.com/locate/msea

Precipitation hardening processes in an


Al0.4%Mg1.3%Si0.25%Fe aluminum alloy
A.K. Gupta a,*, D.J. Lloyd a, S.A. Court b
a

Alcan International Limited, Kingston Research and De6elopment Centre, PO box 8400, Kingston, Ont., Canada K7L 5L9
b
Banbury Laboratory, Southam Road, Banbury, Oxon OX16 7SP, UK
Received 20 June 2000; received in revised form 19 September 2000

Abstract
Aluminum alloys containing Mg and Si as the major solutes are strengthened by precipitation of the metastable precursors of
the equilibrium b (Mg2Si) phase in one or more sequences. There are several metastable particles that can form during ageing, but
the strengthening potential of these particles relative to one another is not extensively studied. In this paper, the precipitation and
hardening potential of different metastable precursors in an Al 0.4%Mg 1.3%Si 0.25%Fe (wt.%) alloy is studied with the help
of differential scanning calorimetry (DSC), transmission electron microscopy (TEM) and hardness measurements. It has been
found that the hardness increase in a freshly solutionized alloy during natural ageing occurs primarily due to precipitation of
clusters and zones, while peak hardness is achieved due to precipitation of b particles. The particles formed following b phase
are not effective hardeners. 2001 Elsevier Science B.V. All rights reserved.
Keywords: Precipitation hardening; Aluminum alloy; Mg2Si

1. Introduction

2. Experimental

Aluminum alloys of the 6XXX series, containing Mg


and Si as the major solutes, are strengthened by precipitation of the metastable precursors of the equilibrium b
(Mg2Si) phase. The precipitation of these metastable
precursors occurs in one or more sequences, which are
quite complex and is not fully understood [1 6]. This is
largely due to confusions regarding the number of
precipitation sequences occurring during ageing, and
also in the number, structure and composition of the
metastable precursors that are involved in a given precipitation sequence [4 8]. Consequently, the strengthening potential of the metastable precursors relative to
one another has not been extensively studied. In this
paper, the precipitation and hardening potential of
different metastable precursors in an Al 0.4%Mg
1.3%Si0.25%Fe (wt.%) alloy is reported.

An alloy containing 0.4% Mg, 1.3% Si and 0.25% Fe


was cast in the laboratory, scalped to remove the
inverse segregation layer, homogenized at 560C, and
then hot and cold rolled to 1 mm thick gauge. Disc
samples of 6 mm diameter were punched from as-rolled
material for differential scanning calorimetry (DSC)
and vickers hardness number (VHN) measurements.
The DSC runs were carried out by placing a 6 mm
disc of the alloy in various heat treated conditions in
the sample pan and a super purity aluminum of equal
mass in the reference pan of the cell. The hardness data
were determined using a 5 kg load from the average of
at least four hardness readings from each sample. The
average hardness data was found to be within a reproducible 3.5 VHN. The cell was equilibrated at 25C and
then heated to 560C at 5, 10, 15 and 20C min 1 in an
argon atmosphere. The heat effects associated with
transformation reactions were obtained by subtracting
a super purity Al baseline run. The results obtained
from the DSC experiments were highly reproducible.
Thin foils for TEM studies were prepared by electropolishing in 30% nitric acid in methanol solution at

* Corresponding author. Tel.: +1-613-5412400; fax: + 1-6135412134.

0921-5093/01/$ - see front matter 2001 Elsevier Science B.V. All rights reserved.
PII: S 0 9 2 1 - 5 0 9 3 ( 0 0 ) 0 1 8 1 4 - 1

A.K. Gupta et al. / Materials Science and Engineering A301 (2001) 140146

27C. The thin foils were examined in a Philips


electron microscope operated at 120 kV. A few of the
samples were also examined in a Hitachi TEM/STEM
operated at 200 kV.
The isothermal ageing curves of the alloy starting
from the freshly solutionized with no natural ageing
and the 3 days naturally aged (T4 Temper) material
were developed at 180C. The non-isothermal ageing
curves were obtained by measuring hardness immediately after heating the as-solutionized or T4 temper
discs in a DSC cell to different temperatures at 10C
min 1.

Fig. 1. Ageing characteristics of an Al0.4%Mg1.3%Si 0.25%Fe


alloy in different ageing conditions. (a) Freshly solutionized and aged
at 180C; (b) freshly solutionized and aged at 200C. (c) T4 temper
and aged at 180C.

Fig. 2. Variation in room temperature hardness of an Al 0.4%Mg


1.3%Si 0.25%Fe alloy due to heating to different temperature at a
heating rate of 10C min 1. Starting temper of the alloy (a) freshly
solutionized; (b) T4 temper.

141

3. Results

3.1. Ageing beha6ior


3.1.1. Isothermal
Fig. 1 shows the hardness curves of the freshly
solutionized and T4 temper alloy obtained after ageing
at elevated temperatures. Curves (a) and (b) in Fig. 1
show ageing curves of the freshly solutionized alloy at
180 and 200C, respectively, while curve (c) shows the
ageing curve of the T4 temper material at 180C. Fig. 1
shows that the hardness of the freshly solutionized alloy
increases rapidly reaching a peak hardness after 8 h of
ageing and then decreases slightly for up to 24 h at
180C. The ageing curve (b) in Fig. 1 obtained for
ageing at 200C shows a peak hardness occurring after
approximately 0.5 h at 200C as opposed to the 8 h it
takes to reach peak hardness when ageing at 180C.
However, the peak hardness is slightly less than that
observed after ageing at 180C. The reduction in peak
hardness value may be related to the reduced matrix
super saturation at high ageing temperatures. The ageing curve (b) in Fig. 1 shows a rapid drop in hardness
beyond peak ageing at 200C.
The ageing curve of the T4 temper material is similar
to that of the as-solutionized material, except that the
hardness values are lower for ageing time ]0.1 h, as
can be seen by comparing curves (a) and (c) in Fig. 1.
3.1.2. Non-isothermal
Curves (a) and (b) in Fig. 2 show the room temperature hardness values of the as-solutionized and T4
temper samples that were non-isothermally heated at
10C min 1 to different temperatures.
It can be seen that the hardness of the freshly solutionized alloy increases relatively slowly with heating to
a temperature of approximately 200C. However, the
hardness then increases rapidly with continued heating
to 250C, and then falls rapidly thereafter with further
heating to approximately 400C. As expected, the peak
hardness obtained in the non-isothermal experiments is
less than that observed in the isothermal ageing experiments as a result of the reduced solute super saturation
of the matrix at high temperatures, (compare Figs. 1
and 2). The material reaches a hardness value lower
than that of the initial hardness upon heating to approximately 305C and beyond. This difference may be
related to depletion of the matrix solute associated with
the formation of coarse Mg2Si during heating.
The alloy in the T4 temper age hardens (curve (b) in
Fig. 2) in a similar manner as that observed in the
as-solutionized material, except for a gradual hardness
decrease up to 210C. The T4 temper material peak
hardens to a slightly lower value than that observed in
the as-solutionized material, although in both cases the
peak hardening occurs at the same temperature.

142

A.K. Gupta et al. / Materials Science and Engineering A301 (2001) 140146

Fig. 3 shows non-isothermal hardness data of the


as-solutionized samples after heat treating to different
temperatures at 5, 10, 15 and 20C min 1. It can be
seen that the heating rate during non isothermal ageing
has a significant influence on the age hardening behavior. For example, the absolute hardness value attained
in the alloy during non-isothermal ageing is decreased
with an increase in the heating rate. In addition, the
extent of the hardness reversion between 220 and 250C
increases with increase in the heating rate (Fig. 3).

3.2. DSC

Fig. 3. Variation in room temperature hardness of a freshly solutionized Al 0.4%Mg 1.3%Si0.25%Fe alloy due to heating to different
temperature at a heating rates (a) 5C min 1; (b) 10C min 1;
(c) 15C min 1; (d) 20C min 1.

Fig. 4. Differential scanning calorimetry (DSC) thermograms of an


Al 0.4%Mg 1.3%Si 0.25%Fe alloy obtained at a heating rate of
10C min 1. Starting temper of the alloy. (a) Freshly solutionized.
(b) T4 temper.

Fig. 5. Differential scanning calorimetry (DSC) thermograms of a


freshly solutionized Al0.4%Mg1.3%Si0.25%Fe alloy obtained by
heating at different rates (a) 5C min 1; (b) 10C min 1; (c) 15C
min 1; (d) 20C min 1.

Fig. 4 shows DSC thermograms of the alloy obtained


at a heating rate of 10C min 1 immediately after
quenching from the solutionizing temperature and also
following three days of natural ageing after solutionizing. The order and number of peaks and troughs in
both curves are quite similar to those observed previously in AA6022 [9] and AA6016 alloys [10]. The
overall shape of the curves is also quite similar to those
observed previously, except for the peak size at 
300C. This difference is probably related to a minor
difference in the composition of the present alloy.

3.2.1. Freshly solutionized material


The curve (a) of the as-quenched material in Fig. 4
shows five precipitation reactions A, B, C, D and E,
and two dissolution troughs F and G. The curves of the
as-solutionized material at four different heating rates,
5, 10, 15 and 20C min 1, are shown in Fig. 5. The
positions of different peaks and troughs in the curves
shift to a higher temperature with increasing heating
rate, suggesting that the precipitation process occurring
during heating are kinetically controlled. In general, the
shifts are proportional to changes in the heating rate,
except within the temperature regime of the dissolution
troughs F and G where multiple precipitation and
dissolution reactions are occurring simultaneously. The
net heat effect associated with these reactions gives rise
to complex dissolution troughs F and G containing
peak E. The absence of a plateau between the different
reactions is expected in situations where the change in
solvus composition is highly dependent on the temperature, one particle type is transforming to another and/
or two or more independent reactions are occurring
simultaneously.
3.2.1.1. Isothermally aged freshly solutionized material.
Fig. 6 shows the DSC curves obtained from the as-solutionized materials that were aged for different times at
180C prior to heating in the DSC cell at 10C min 1.
It can be seen that ageing for 30 min completes the
precipitation within peak A, almost finishes the process
within peak B, and induces a small amount of precipitation due to peak C (compare curves (a) and (b) in

A.K. Gupta et al. / Materials Science and Engineering A301 (2001) 140146

Fig. 6). The dissolution trough F is partially replaced


by a plateau up to 190C which is then followed by a
broad composite peak marked H. It appears that the
precipitation process D is occurring at slightly lower
temperature, causing the peak to overlap with peak C
to form a broad peak H in curve (b) of Fig. 6. In
addition, the precipitation peak E within the dissolution
trough G is reduced to a minor peak E.
The DSC curves obtained from samples aged for
longer times show features quite similar to curve (b) of
Fig. 6. The only difference is that the reaction rate
(peak height) and the extent of precipitation (peak area)
under the peak H increases with an increase in ageing
times up to 8 h and then decreases upon further ageing
to 24 h.
The peak hardness value is achieved after 8 h at
180C. The increase in height of peak C with ageing
and the presence of a dissolution trough F appears to
suggest that the precipitation process C is proceeding at

Fig. 6. Differential scanning calorimetry (DSC) thermograms of a


freshly solutionized Al0.4%Mg1.3%Si0.25%Fe alloy after ageing
at 180C for different times. The curves have been offset for clarity.
(a) Freshly solutionized; (b) 30 min; (c) 1 h; (d) 4 h; (e) 8 h; (f) 24 h.

Fig. 7. Differential scanning calorimetry (DSC) thermograms of a


freshly solutionized Al0.4%Mg1.3%Si0.25%Fe alloy after heating to different temperatures in a DSC cell at 10C min 1. The
curves have been offset for clarity. (a) Freshly solutionized; (b) 151C;
(c) 210C; (d) 234C; (e) 250C; (f) 305C.

143

the expense of precipitates that are formed within peak


B. The precipitation of peak C reaches a maximum
after 8 h and then slows down upon further ageing. The
presence of trough F in curve (a) of the as-solutionized
sample and the absence of the same trough in the traces
of the aged samples (Fig. 6) suggests that the processes
associated with B, F and C peaks/troughs are overlapping. Peak C is also overlapping with the high temperature peak D.

3.2.1.2. Non-isothermally aged material. Fig. 7 shows


the DSC curves obtained from the as-solutionized samples that were heated in a DSC cell to different temperatures, cooled to room temperature and then reheated
to 560C at 10C min 1. Peak A present in curve (a) of
Fig. 7 is absent in curve (b), suggesting that precipitation within peak A is complete after heating up to
 150C. Trough F disappears by heating to 234C,
with no significant changes taking place in the shape of
peak D, compare curves (a), (b) and (c) in Fig. 7. Curve
(d) shows that the trough F is replaced by a minor
broad peak B and minor dissolution trough F after
heating to 250C. The appearance of peak B (curve (d)
in Fig. 7) seems to suggest that peak B is possibly a net
effect of precipitation and dissolution reactions occurring between 150 and 250C. The presence of trough F
in curve (d) of Fig. 7, suggests that the particles from
peak B are dissolving prior to precipitation giving the
following peak C. Based on these observations, it can
be said that the precipitates forming within peak B do
not transform continuously into particles giving rise to
peaks C and D in curve (a) of Fig. 7. These results are
consistent with the curves obtained from the isothermally aged samples, compare Figs. 6 and 7.
3.3. Transmission electron microscopy (TEM)
3.3.1. Isothermally aged as-solutionized alloy
Fig. 8(a) shows a bright field (BF) TEM image
representing the typical microstructure of the as-solutionized material aged 30 min at 180C. The image
shows small needle shaped b particles aligned in [100]
matrix direction and homogeneously distributed
throughout the matrix. The presence of a large number
of tiny dots represents the b needles viewed end-on.
The microstructure is quite similar to that reported by
Perovic et al. [7] on an AA6016 alloy.
Fig. 8(b) and (c) show the typical microstructure of a
sample aged 8 h at 180C. The bneedles of 4 nm
diameter and  40 nm length are similar to those
observed in Fig. 8(a), except they are larger in size. Fig.
8(c) shows the microstructure after tilting 510 from
the [100] zone axis. The particles were analyzed by
energy dispersive X-ray spectroscopy (EDS) and revealed only a Si peak.

144

A.K. Gupta et al. / Materials Science and Engineering A301 (2001) 140146

Fig. 8. Bright field images of a freshly solutionized Al0.4%Mg 1.3%Si 0.25%Fe alloy after ageing at 180C for different times (a) 30 min; (b
and c) 8 h; (d) 24 h.

Fig. 8(d) shows the typical micrograph of the sample


aged 24 h at 180C. The b needles observed in Fig.
8(b) have now been replaced by long rods and irregularly shaped particles. Most of the long rods are
bphase, while a few rods and irregularly shaped particles are again due to Si.

3.3.2. Non-isothermally aged material


The microstructural changes that occur during heating up to different temperatures are indicated on the
DSC and hardness curves in Fig. 9. Fig. 9(a) shows the
BF image of a sample after heating to 234C. This
image shows  4 nm diameter and 15 20 nm long
needles aligned along 100 matrix direction and tiny
dot shaped b particles responsible for the occurrence
of peak B and the hardening, Fig. 9.
Fig. 9(b) shows the BF image of a peak hardened
sample heated to 250C. The needle shaped particles
seen in Fig. 9(a) are now 40 nm in length. The
microstructure also contains tiny dots of b needle
viewed end on, long elongated rods and irregular particles on dislocations, marked A, B, and C in Fig. 9(b),
respectively. The presence of the rods B and the precipitate free zone around them suggests dissolution of

some fine coherent needles and the formation of a more


stable precursor of the equilibrium phase. The BF
image of a sample heated to 305C is shown in Fig.
9(c). The needles are coarser than those seen in Fig.
9(b) and new particles marked 1 and 2 appear in the
structure. The particles marked 18 were chemically
analysed using a STEM and all of them found to
contain Mg and Si.
4. Discussion
The decomposition of the supersaturated solid solution (SSS) of AlMgSi alloys containing excess Si is
believed to proceed in the following sequence [412].
SSS (Mg+ Si)clusters/GP(I)platelike/spherical
b/GP(II)needles brods + Si+others
bplates + Si
The decomposition process begins with the formation
of two types of (Mg+ Si) clusters or zones [2,8]. The
first to form is coherent Si rich (Mg+ Si) cluster, which
is then enriched by diffusion of Mg atoms to form the
Si-depleted platelike [6] or spherical [7] (Mg+Si) clusters or GP(I) zones. As ageing proceeds, the clusters

A.K. Gupta et al. / Materials Science and Engineering A301 (2001) 140146

and zones become ordered and develop the needle


shaped GP(II)/b phase. There is considerable disagreement regarding the number, structure and composition
of the metastable phases and their effect on hardening
following the b formation [7]. Upon prolonged ageing,
the formation of b and B with Mg/Si ratio between 0.9
and 1.2 [4,13] together with Si [2] has been suggested.
Matsuda et al. have suggested the precipitation of three
independent phases, namely A, B and C, in addition to
the b phase [3,5]. The atomic ratio of Mg to Si in the
b phase is about Mg/Si 1.68, while Type A and Type
B contain three elements Al:Mg:Si in ratios of 4:1:5 and
4:2:5, respectively [5]. The precipitation of b and Sirich Type B phases occurs simultaneously, which is
then followed by formation of Si rich Type A particles.
Following complete dissolution of the b particles, the
Type B starts dissolving and further increases precipitation of Type A particles. The Type C particle is equivalent to B [5].
The clusters and zones are too difficult to resolve by
conventional TEM, however their formation is associated with the hardness increase from  50 VHN to
 72 VHN as seen in Fig. 1. Fig. 8(a) shows that most
of the clusters and zones are replaced by needle shaped
coherent GP(II)/b upon ageing 30 min at 180C. At
this stage, the peak B in DSC curve (b) of Fig. 6 almost

145

disappears and hardness is increased from 72 to


 104 VHN, as seen from curve (a) of Fig. 1. The
presence of a hump B in DSC curves of Fig. 6 suggests
that the b precipitation continues to occur up to 8 h at
180C. The peak aged structure contains mainly coherent b and a few Si particles (Fig. 8(b)). The freshly
solutionized material acquires  115 VHN hardness in
the peak aged condition. The material begins to lose
strength with further ageing beyond peak. The hardness
loss is very rapid during overageing at higher ageing
temperatures, curve (b) of Fig. 1. The microstructure in
the over aged condition mainly contains b particles
along with Si.
The non-isothermal ageing curves in Fig. 2 also
suggest that the peak hardening is mainly due to the
presence of the b phase. This is supported by the fact
that unlike the DSC curves, the non-isothermal ageing
curves (Fig. 2) show a single hardness peak that nearly
coincides with the b peak B in Fig. 9. Although no
attempt was made to identify precipitates B and C, it is
clear that these particles do not provide hardening.
Generally, the particles that are formed during later
stages of ageing, such as B, Type A, Type B and Type
C reported in the literature [35,14], are not very
effective hardeners. The peak hardening in the asquenched alloy occurs primarily due to the b phase,

Fig. 9. Variations in microstructure and hardness of a freshly solutionized Al 0.4%Mg 1.3%Si 0.25%Fe alloy due to heating at 10C min 1 to
different temperature. The particles marked 18 were chemically analysed in a STEM. (a) Heated to 234C; (b) heated to 250C; (c) heated to
305C.

146

A.K. Gupta et al. / Materials Science and Engineering A301 (2001) 140146

which is consistent with the results of recent studies


[7,9].
The hardness peaks observed in non-isothermal hardness curves of Fig. 3 are associated with peaks A and B.
A small decrease in hardness (Fig. 3) between 150 and
215C is related to the dissolution process associated
with trough F in curve (a) of Fig. 7. The high temperature precipitation reactions due to peaks C and D
(curve (a) in Fig. 7) do not provide hardening.
It should be noted that the hardening is related to the
heating rate during non-isothermal ageing (Fig. 3). In
general, the hardness value obtained at 5C min 1 is
higher than those obtained at faster heating rates. In
addition, the hardness peak shifts to higher temperatures and the peak hardness value reduces to a lower
value with increasing heating rates. The latter may
suggest that the b solvus is strongly dependent on the
alloy composition at high temperatures, which would
reduce the matrix super saturation due to increased b
reversion with increasing heating rate. The reduction in
the peak hardness values will also occur due to coarsening of the b and formation of particles marked B and
C due to peak shifts to high temperatures at higher
heating rates.
The precipitation process in the T4 temper material
appears similar to its as-quenched counterpart, except
that hardening in the early stage of ageing (30 min at
180C) is significantly lower than its as-quenched equivalent. The extent of this loss is however partially recovered due to precipitation of the b particles upon
further ageing. This loss is believed to be due to the
reversion of very fine zones and clusters that are formed
during natural ageing [15].

5. Conclusions
The hardness increase during natural ageing of an
Al0.4%Mg1.3%Si 0.25%Fe aluminum alloy occurs
primarily due to precipitation of clusters and zones.
The choice of heating rate used to achieve the ageing
temperature has a significant effect on the hardness
reversion during ageing of the freshly solutionized
material.
The peak hardness in the as-quenched AA6016 variant is primarily caused by precipitation of b phase.

Natural ageing following the solutionizing treatment


reduces the age hardenability of the Al0.4%Mg
1.3%Si0.25%Fe alloy, especially in the under aged
condition.
The microstructure of the freshly solutionized alloy
shows at least two additional phases B and C occurring
in the overaged condition. These phases together with
those formed following the b phase are not effective
hardeners.
Acknowledgements
The authors would like to thank Alcan International
Limited for permission to publish this work and to Ms
J.M. Bookbinder for technical assistance.
References
[1] I. Dutta, S.M. Allen, J. Mater. Sci. Lett. 10 (1991) 323.
[2] A.K. Gupta, D.J. Lloyd, in: L. Arnberg, et al. (Eds.), Aluminum
Alloys: Their Physical and Mechanical Properties, vol. 2, Norwegian Institute of Technology and SINTEF Metallurgy, Trondheim, 1992, p. 21.
[3] K. Matsuda, Y. Uteni, H. Anada, S. Tada, S. Ikeno, in: L.
Arnberg, et al. (Eds.), Aluminum Alloys: Their Physical and
Mechanical Properties, vol. 1, Norwegian Institute of Technology and SINTEF Metallurgy, Trondheim, 1992, p. 272.
[4] G.A. Edwards, K. Stiller, G.L. Dunlop, M.J. Couper, Acta
Mater. 46 (11) (1998) 3893.
[5] K. Matsuda, Y. Sakaguchi, Y. Miyata, Y. Uteni, T. Sato, A.
Kamio, S. Ikeno, J. Mater. Sci. 35 (2000) 179.
[6] K. Matsuda, H. Gamada, K. Fuji, Y. Uteni, T. Sato, A. Kamis,
S. Ikeno, Metall. Mater. Trans. 29A (1998) 1161.
[7] A. Perovic, D.D. Perovic, G.C. Weatherly, D.J. Lloyd, Scripta
Mater. 41 (7) (1999) 703.
[8] A.K. Gupta, D.J. Lloyd, Met. Trans. 30A (1999) 879.
[9] M.F. Miao, D.E. Laughlin, Scripta Mater. 40 (7) (1999) 873.
[10] A.K. Gupta and D.J. Lloyd, Proceedings of 34th Annual Conference on Recent Metallurgical Advances in Light Metal Industries, In: S. MacEwen, J.-P. Galardeau (Eds.), The Metallurgical
Society of the Canadian Institute of Mining, Metallurgy and
Petroleum, Vancouver, British Columbia, August 2024, 1995,
p. 243.
[11] N. Muruyama, R. Uemori, N. Hashimoto, M. Saga, M.
Kikuchi, Scripta Mater. 36 (1) (1997) 89.
[12] J.P. Lynch, L.M. Brown, M.H. Jacobs, Acta Metall. 30 (1982)
1389.
[13] M. Murayama, K. Hono, Acta Mater. 47 (5) (1999) 1537.
[14] S.D. Dumolt, D.E. Laughlin, J.C. Williams, Scripta Met. 18
(1984) 1347.
[15] D.W. Pashley, J.W. Rhodes, A. Sendorek, J. Inst. Met. 94 (1966)
41.

You might also like