You are on page 1of 11

Materials and Design 181 (2019) 107927

Contents lists available at ScienceDirect

Materials and Design

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

Age-hardening of high pressure die casting AlMg alloys with Zn and


combined Zn and Cu additions
Lukas Stemper a,⁎, Bernhard Mitas b, Thomas Kremmer b, Steffen Otterbach d,
Peter J. Uggowitzer b,c, Stefan Pogatscher a,b,⁎⁎
a
Christian Doppler Laboratory for Advanced Aluminum Alloys, Chair of Nonferrous Metallurgy, Montanuniversitaet Leoben, Franz-Josef Straße 18, 8700 Leoben, Austria
b
Chair of Nonferrous Metallurgy, Montanuniversitaet Leoben, Franz-Josef Straße 18, 8700 Leoben, Austria
c
Laboratory of Metal Physics and Technology, Department of Materials, ETH Zurich, 8093 Zurich, Switzerland
d
Audi AG, N/PG-T52, 74148 Neckarsulm, Germany

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• New high pressure die casting alloys


based on Al-Mg-Zn-(Cu) for structural
cast components are studied.
• Two-step artificial aging can either en-
hance the maximum hardness or accel-
erate the hardening kinetics strongly.
• Heat treatment strategy and alloy
chemistry significantly influence the
microstructural evolution during heat
treatment.

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

Article history: This study investigates the age-hardening of AlMg alloys with Zn and combined Zn and Cu additions. Two
Received 9 April 2019 AlMg5Mn1 alloys modified with Zn and Cu were processed by high pressure die casting (HPDC) and different
Received in revised form 30 April 2019 heat treatment strategies. Single step artificial aging, artificial aging with pre-aging and the effect of the
Accepted 8 June 2019
quenching rate were studied via hardness measurements and transmission electron microscopy (TEM). Single-
Available online 9 June 2019
step artificial aging resulted in an increase in hardness of 58% in peak aged condition for the Zn-only modified
Keywords:
alloy and of 56% for the Zn- and Cu-containing alloy. Pre-aging treatments either reduce the necessary aging
Aluminum alloys time or increase the hardness, depending on the parameters used. Microstructural investigations indicate a sig-
AlMg nificant change in the S- or T-phase precursors, and in precipitation density with pre-aging. The alloys have high
Alloy design potential for use as complex structural HPDC components in lightweight transport applications, but are also of
Hardening behavior general interest for components which require high strength and formability.
Precipitation © 2019 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://
High pressure die casting creativecommons.org/licenses/by-nc-nd/4.0/).

⁎ Corresponding author.
⁎⁎ Correspondence to: S. Pogatscher, Christian Doppler Laboratory for Advanced Aluminum Alloys, Chair of Nonferrous Metallurgy, Montanuniversitaet Leoben, Franz-Josef Straße 18,
8700 Leoben, Austria.
E-mail addresses: lukas.stemper@unileoben.ac.at (L. Stemper), stefan.pogatscher@unileoben.ac.at (S. Pogatscher).

https://doi.org/10.1016/j.matdes.2019.107927
0264-1275/© 2019 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
2 L. Stemper et al. / Materials and Design 181 (2019) 107927

1. Introduction hardening effect (similar to that of AlCuMg alloys) at early stages of ar-
tificial aging at higher temperatures (180 °C), related the strength in-
One of the most significant changes in people's lives over the last few crease of approx. 30 MPa to the formation of S″-Phase, which is also
decades has involved the growth of transport and traffic. While environ- the major hardening component in 2xxx-series alloys [12]. The peak
mental and sustainable aspects had only minor relevance in earlier strength was found to occur after 11 days. Further investigations by
years, they have assumed political and economic importance due to Ratchev [13] focused on the influence of the Cu/Mg ratio (AlMg4.2Cu0.6,
the rising CO2 emissions associated with increasing traffic [1,2]. A AlMg3Cu0.6, AlMg3Cu0.9) on age hardening behavior. While high Cu/
well-established approach to these challenges in the automotive indus- Mg ratios generate an improvement in the age hardening response
try is mass reduction via light-weight construction and materials de- over the whole aging period, low ratios cause a delay in the precipitation
sign. Besides the substitution of heavy materials such as steel with sequence and a minor response. Reducing the aging temperature from
light aluminum alloys, a major aspect of light-weighting is the creation 180 °C to 140 °C and a lower Mg content only had a minor effect on
of complex structural components which combine the individual pur- the strength increase in the early stages.
poses of several parts in one. Although the price of such components Besides copper, the addition of Zn to AlMg alloys has also been well
is higher, total costs decrease due to diminished joining effort [3]. Typi- reported. In several cases these alloys were treated as 7xxx-series alloys
cal examples can be found in car body applications such as strut hous- with very small Zn/Mg ratios (b 1). Age hardening is generated by pre-
ings or side and cross members. The complex shapes of these, cipitation of T-phase (Mg32(Al,Zn)49) and its precursors. The exact de-
resulting from varying wall thicknesses and transition radii, makes scription of the sequence is still disputed, but the difference in crystal
them ideal candidates for high pressure die casting (HPDC) in mass pro- structure and chemical composition between hardening precursors
duction [4]. and equilibrium phase is expected to be only minor [14]. EN AW-7xxx
Several standardized alloys are available which have been optimized hardened by T-phase shows higher strength and increased resistance
for HPDC processing. Due to their good producibility, cast alloys contain- against intergranular corrosion (IGC) compared to alloys with the
ing silicon (EN AC-4xxxx) close to the eutectic composition dominate this usual Zn/Mg ratios hardened by the η-phase precipitation sequence
field [5]. A major disadvantage of this alloy class, however, is limited [15]. These beneficial effects are also observed in Zn-modified AlMg al-
strength and formability, which limits its use in vehicle body panels. In loys, which exhibit increased corrosion resistance (especially IGC) com-
particular, lack of ductility makes remedial action necessary to enable pared to sensitized 5xxx-series alloys due to the suppression/
punch riveting operations. Even though AlSi10MnMg, a highly optimized modification of the unfavorable β-phase [16–19] and increased strength
and very commonly used alloy in crash-relevant applications, offers high caused by precipitation hardening [16,20,21]. Interestingly, serrated
strength due to its age hardenability combined with acceptable ductility, flow due to dynamic strain aging, which is a well-known phenomenon
some efforts have been made to develop and improve AlMg cast alloys in 5xxx-series alloys and a result of interactions between solute Mg
(EN AC-5xxxx). These alloys show a quite attractive combination of mod- atoms and dislocations during deformation [22], is reported to be par-
erate strength and increased ductility, including an improvement in cor- tially suppressed by adding Zn; this occurs via formation of Zn-Mg clus-
rosion resistance according to their high Mg content [5,6]. On the other ters during natural aging and precipitate formation during artificial
hand, their wide solidification interval makes the casting operation aging, which causes depletion of solute Mg in the matrix [20,23].
more difficult and limits the design flexibility of the cast products. The improved corrosion resistance in both the Cu- and Zn-modified
At present, replacing AlSi cast alloys with AlMg cast alloys generates AlMg alloys led to in-depth investigations on the effect of combined ad-
only minor benefits in final product properties, and is accompanied by dition of these elements. It was found that certain amounts of Cu can
deterioration in processing. Therefore, effort is required to provide balance the impairing effect of Zn on SCC and even improve the me-
these alloys with better mechanical characteristics. In AlMg wrought al- chanical properties strength and work hardenability [24–26]. Recently
loys (EN AW-5xxx), which exhibit the best formability of all aluminum this kind of modification, with increased alloy content (up to 3 wt% Zn
alloy classes due to their high strain hardening assets (resulting from and 0.45 wt% Cu), has attracted attention due to the significant
the high content of Mg in solid solution), several approaches to increas- strengthening effect by precipitation hardening generated. While Zn-
ing their only moderate strength have been investigated [7]. Although modified AlMg alloys exhibit a relatively high hardening response
the binary system of Al and Mg shows a thermodynamic tendency of resulting from T-phase and its precursors only after a certain incubation
precipitate formation, non-beneficial kinetics impede a controllable period, Cu-modified alloys show a rapid hardening effect at early stages
and favorable use of precipitation hardening by β-Phase (Al3Mg2) and of aging due to GPB-zone formation, but reach moderate peak hardness
its precursors, which renders this alloy class non-heat-treatable [4]. In resulting from S-phase and its precursors only after longer aging times.
addition to processing optimizations and alloy development within The combined addition of both elements improves the age-hardening
the standardized ranges, the deployment and development of precipita- response and leads to higher peak hardness at shorter aging times
tion hardening by major alloy modifications with zinc (Zn) and copper through the synergistic effects of T- and S-phase precipitates [27]. The
(Cu) have been found to significantly improve the mechanical application of a two-stage artificial aging procedure generates an al-
properties. tered precipitation microstructure and an improved age-hardening re-
A very early approach by Hino [8] focused on avoiding softening dur- sponse compared to single-stage treatment. Peak hardness is caused
ing paint baking (175 °C/30 min) by adding 0.5 wt% Cu (KS5030, by fine and equiaxed precipitates of Cu-modified T-phase [28]. Investi-
KS5032). Precipitation of S′-Phase (Al2MgCu) was found to balance gations on the IGC resistance of these alloys have revealed precipitation
the softening generated by recovery of the 2% pre-deformed sheets. free zones (PFZ) at the grain boundaries for alloys containing only Zn,
Benefits for formability, stress corrosion cracking (SCC) and filiform cor- but also for those alloyed with Zn and Cu. Cu-containing alloys showed
rosion have also been reported. More recent investigations on Cu- the effects of narrower PFZs and higher density of precipitates in the
modified alloys (EN AW-5182, EN AW-5083) and Cu-containing alloys grains, which are responsible for higher resistance to IGC [29]. The pre-
(AA 5023) confirm these results. Depending on the thermo- cipitation sequence for this kind of alloy seems to depend on the aging
mechanical treatment, precipitation hardening by S-phase and its pre- strategy. While at single-stage artificial aging two precipitation mecha-
cursors can reduce [9] or even compensate [10] for the decrease in nisms compete (SSSS (super-saturated solid solution) → GPB-zones → S
strength caused by paint bake treatment. Long-term natural aging, ″ → S′ → S-phase (Al2MgCu) for the Cu-containing phase and SSSS →
which can be a crucial factor for age hardenable alloys like the 6xxx- GPI-zones → GPII-zones (T") → T' → T-phase (Mg32(Al, Zn)49) for the
series, showed only a slight increase in strength of only few MPa with- Zn-containing phase), two-stage aging follows the sequence of T-
out any significant effect on ductility in AA 5023 [9]. Studies on phase precipitation with the copper atoms incorporated in the T-
AlMg4.2Cu0.6 by Ratchev [11], which concentrated on the rapid phase (Mg32(Al, Zn, Cu)49) [27,28,30].
L. Stemper et al. / Materials and Design 181 (2019) 107927 3

Table 1 temperature range of −10 °C to −20 °C and a voltage range of 10 V to


Main element content of the investigated alloys in wt%. 20 V. High angle annular dark field (HAADF), bright field (BF) and dif-
Mg Mn Zn Cu fractometry studies were carried out using a FEI Tecnai-F20 HRTEM.
Alloy 1 5.0 0.8 3.8 0.0
Alloy 2 4.8 0.8 3.4 0.5 3. Results

3.1. Single-step artificial aging at varying temperatures


Taking all this into account, it seems reasonable to determine the ap-
plicability of these measures on cast alloys. If such measures led to sig- The hardness change of alloy 1 and alloy 2 during single-step aging
nificant improvements, the trade-off between product properties and at 125 °C (red lines) and 175 °C (blue dotted lines) are pictured in
processing would change in favor of AlMg alloys and ultimately result Fig. 2a and b, respectively. While no substantial hardening occurs at
in a better product. This paper illustrates the results of a systematic 175 °C, a remarkable increase in hardness can be observed at the
study on the age hardenability of Zn- and/or Cu-modified EN AW- lower aging temperature of 125 °C for both alloys. The Cu-containing
5083, manufactured by HPDC. The influence of varying both alloying alloy 2 also exhibits an early aging response after a few minutes and
content and behavior in different heat treatment strategies is a key as- gains more hardness over the investigated time range than the Zn-
pect of this study. In addition, the microstructure was analyzed by only modified alloy 1. The onset of hardness increase for alloy 1 is also
TEM investigations to determine the main hardening phases and their shifted to longer aging time compared to alloy 2. Peak hardness was
distribution within the grain structure. not reached over the aging time studied. Contour-plots indicating the
measured hardness at certain time/temperature-pairings for both alloys
2. Experimental are shown in Fig. 2c and d. It is apparent that temperatures above 150 °C
have hardly any effect on hardness, but that decreasing aging tempera-
Table 1 presents the main element content of the alloys investigated. ture down to 100 °C accelerates and enhances the hardening response.
Both are based on EN AW-5083 supplied by AMAG casting GmbH and Further temperature reduction leads to re-declination of the hardening
modified by adding varying amounts of Zn and Cu. rate but still results in high hardness values after longer aging times.
The alloys were produced at the industrial scale and shed to 3 mm This effect is more distinct in the Cu-containing alloy 2. Again, peak
thick plates with a format of 120 mm × 100 mm by HPDC. Samples hardness was not reached at any temperature over the treatment
were taken randomly over the casting period and further processed to duration.
hardness samples by cutting and grinding.
A circulating air furnace (Nabertherm N15/65 SHA) was used to per- 3.2. Influence of pre-aging treatments on age-hardening response
form solution annealing at a temperature of 465 °C for 35 min. To gen-
erate a supersaturated solid solution, the samples were quenched by The influence of pre-aging was investigated with an eye on two
immersion in water at room temperature or, for the investigation of major aspects: pre-aging duration at 100 °C (according to the highest
quench sensitivity, cooled by air at standard conditions on a wire grid. hardening rate in single-step aging) and second-step artificial aging
While pre-aging and artificial aging at higher temperatures was per- temperature after the most beneficial pre-aging treatment. Hardness
formed in a circulating oil bath, natural aging took place at room tem- curves for several pre-aging times are plotted in Fig. 3 for alloy 1
perature. A full list of all heat treatments investigated, including a (a) and alloy 2 (b). Compared to single-step aging at 150 °C, all pre-
schematic representation, is shown in Fig. 1. aging treatments (15 min, 60 min and 100 min at 100 °C) show in-
All pictured hardness values in this study represent an average of creased hardness at the start but a decrease in strength over the first
five independent hardness measurements which were taken by an couple of minutes of the second aging step, which is more distinct in
EMCO-TEST M4 unit according to Brinell's method (HBW 2.5/62.5). the Cu-free alloy 1. While longer pre-aging resulted in an accelerated
Thin-foils for transmission electron microscopy (TEM) were pre- and increased hardening response in this alloy, only a minor difference
pared by grinding hardness samples to a thickness of approx. 100 μm. could be observed when varying pre-aging times in the Cu-containing
Punched out 3-mm discs were further processed by twin jet electro- alloy 2, even though the level of hardness is higher compared to the
polishing using a solution of 75% methanol and 25% nitric acid at a hardness profile without pre-aging treatment. It is also apparent that

Fig. 1. Schematic representation of the heat treatments and applied parameters.


4 L. Stemper et al. / Materials and Design 181 (2019) 107927

increased pre-aging time generates enhanced maximum hardness. The longer natural aging times. An applied pre-aging treatment (100 °C/3 h)
influence of the final aging temperature after 3 h of pre-aging at 100 °C before natural aging (blue dotted line) generates a higher initial strength
is plotted in Fig. 3c (alloy 1) and Fig. 3d (alloy 2). Higher temperatures in level (especially for alloy 2), but also provides increased stability over at
the second aging step clearly shift the hardness increase to shorter least 6 days of aging for the Zn-only-modified alloy 1, and over 24 days
times, but also limit the maximum achievable hardness. Pre-aging treat- for alloy 2. After 175 days at room temperature both the pre-aged and
ment in combination with increased temperature in the final aging step the non-pre-aged samples exhibit hardness values of 128 HBW and
also results in over-aging within the observed time interval for both al- more, which indicates an undesired strength increase. On the other
loys. Again, the general hardness level is higher in alloy 2 but a slight hand, two weeks of natural aging after solution heat treatment and
softening is visible at early aging times. quenching produces the highest values measured during subsequent arti-
ficial aging of all conditions investigated. The hardness profiles for this
3.3. Natural aging and its effect on age-hardening response heat treatment strategy are pictured in Fig. 4c and d (blue dotted lines)
and compared to single-step aging curves at 125 °C (red line). During
Fig. 4a and b (red lines) shows the evaluation of the stability of the so- the first 30 s of aging the hardness level resulting from natural aging
lution treated and quenched condition during natural aging. A clear in- stays constant, before dropping almost to the level gained at single-step
crease in strength is apparent for both alloys, and Cu alloying seems to aging. With progressive aging time the profile rises again to the highest
delay the onset to a certain degree but results in a higher hardness at values, of 160 HBW for alloy 1 and even 170 HBW for alloy 2.

Fig. 2. Single-step artificial aging of alloy 1 (a) and alloy 2 (b) at 125 °C (red line) and 175 °C (blue dotted line); contour plots of single-step artificial aged alloy 1 (c) and alloy 2 (d) over a
temperature range of 75 °C to 200 °C.
L. Stemper et al. / Materials and Design 181 (2019) 107927 5

3.4. Effect of the quenching rate on age-hardenability a narrower width of approx. 200 nm (Fig. 6b). Within the grain the pre-
cipitates show a fine, uniform distribution for both alloys (Fig. 6e and f),
In Fig. 5 single-step artificial aging hardness curves (100 °C), but they are little coarser and more elongated in the Cu-free alloy 1
resulting from water quenching (red line) and air cooling (blue dotted (Fig. 6e). This might explain its slightly lower hardness (compare red
line) after solution heat treatment at 465 °C, are plotted for alloy 1 lines, Fig. 2a and b).
(a) and alloy 2 (b). Both alloys exhibit noticeable quench sensitivity, Two-step artificial aging (100 °C/3 h + 175 °C/3 h), which repre-
causing significant lower hardness with air cooling applied. This effect sents a fast and thus beneficial age hardening response, generates sig-
is more distinct for alloy 2. nificantly narrower PFZs with widths of approx. 200 nm for alloy 1
(Fig. 6c) and far b100 nm for the Cu-containing alloy 2 (Fig. 6d). It
3.5. Microstructural investigations also generates an increased amount of grain boundary precipitates
(GBPs), as seen in Fig. 6c and d. Compared to single-step aging, two-
The microstructure of both alloys studied was examined with re- step heat treatment produces much smaller precipitates, which appear
spect to aspects of phase distribution (Fig. 6) and phase identification in a drastically increased number density of precipitates. Again, the Cu-
(Fig. 7), with a focus on the influence of the heat treatment strategy. containing alloy 2 expresses this effect even more strongly, generating
For single-step artificial aging (125 °C/48 h), which represents the al- an explicitly higher hardness level (compare orange dashed line in
most peak aged condition, both alloys show clear appearance of PFZs Fig. 3c and d).
and precipitates at grain boundaries. For the Zn-only modified alloy 1 The diffraction study of both alloys investigated shows a difference
the width of the PFZ exceeds 500 nm (Fig. 6a), but Cu addition (alloy in the presence of hardening phases. A comparison between simulated
2) produces a decreased expression of this phenomenon and results in diffraction patterns for the phases expected (predicted by Pandat [31],

Fig. 3. Two-step artificial aging of alloy 1 (a, c) and alloy 2 (b, d) with varying pre-aging time (a, b) and varying artificial aging temperature (c, d).
6 L. Stemper et al. / Materials and Design 181 (2019) 107927

S-phase: Al2MgCu, orthorhombic, aS = 0.400 nm, bS = 0.923 nm, cS = (Fig. 7d). This suggests that the precipitation process differs in this alloy
0.714 nm [11] and T-phase: Mg32(Al,Zn)49, cubic, aT = 1.416 nm, bT = according to whether one-step or two-step aging is applied.
1.416 nm, cT = 1.416 nm [32,33]) and the experimental data is provided
in Fig. 7. Due to their similar crystal structures a distinction between 4. Discussion
precursors and equilibrium precipitation phases was not included in
this study [14,34–36]. Due to the wealth of data in the literature (exper- The results presented in the previous chapter can be summarized by
imental and theoretical) the experimental results depicted were ac- the following observations:
quired with the electron beam direction of 〈001〉 to facilitate
evaluation of the diffraction pattern. The alloy without Cu-addition (i) For single-step aging, it has been found that temperatures above
(alloy 1) exhibits only spots corresponding to T-phase or its precursors 150 °C only cause a slight hardness increase, while lower temper-
at the 2/5 and 3/5 〈022〉Al position (Fig. 7b), after one-step and two-step atures significantly increase the hardening response; this is even
aging [28]. more pronounced for the Cu-containing alloy 2. Temperatures
Due to the Cu-addition in alloy 2, the formation of a Cu-rich S-phase below 100 °C still lead to increased hardness but the onset is
or its precursors is possible. This is indeed observed after single-step ar- shifted to longer aging times.
tificial aging treatment. Fig. 7c clearly exhibits additional reflexes at the (ii) Pre-aging treatments indicate that such procedures can offer su-
predicted position (Fig. 7a) for S-phase between the {002}Al and {022}Al perior hardness and kinetics. Increased pre-aging time generates
spots. During the two-step artificial aging process, no S-phase precur- a higher, accelerated hardness increase. In combination with ele-
sors are formed, as can be seen in the corresponding diffraction pattern vated temperatures in the second aging stage, even greater

Fig. 4. Natural aging of quenched and quenched plus pre-aged alloy 1 (a) and alloy 2 (b). The influence of prior natural aging on subsequent artificial aging at 125 °C of alloy 1 and alloy 2 is
shown in (c) and (d).
L. Stemper et al. / Materials and Design 181 (2019) 107927 7

acceleration of the hardening response is possible, but unfortu- and thus to a higher activation energy and larger critical nucleus for
nately this is accompanied by a decreased maximum hardness the competing T-phase precursor formation. Once the energy barrier
level. is overcome, precipitation of the Zn-rich T-phase precursors accelerates
(iii) Natural aging after solution heat treatment indicates poor hard- due to the higher diffusion rate of Zn atoms. With consumption of Mg
ness stability at room temperature storage which, however, is atoms, the Cu/Mg ratio increases and S-phase precursor formation is fa-
improved by pre-aging treatment. On the other hand, aging at vored again. Peak hardness is gained by synergetic effects of both
125 °C after two weeks of prior natural aging generated the phases.
highest hardness values of approx. 160 HBW and 170 HBW for For the lower aging temperature (125 °C) the age hardening re-
alloy 1 and alloy 2, respectively. sponse in both alloys is much more pronounced, particularly at longer
(iv) The hardening response is significantly influenced by the occur- aging times. For alloy 1 this effect is judged to result from the higher
rence of quench sensitivity. precipitation pressure caused by higher Mg-supersaturation at lower
temperatures. Provided that Mg is completely in solid solution, i.e. no
solute loss occurs during quenching, thermodynamic calculations
using Pandat [31] predict a Mg-supersaturation of 3.77 wt% at 125 °C
We discuss these findings in detail below. and 3.2 wt% at 175 °C, respectively. The roughly 13% higher supersatu-
ration at the lower aging temperature is assumed to cause both faster
4.1. Single-step artificial aging precipitation kinetics of T-phase precursors and an enhanced hardening
response.
According to studies of precipitation behavior in Zn-modified [37] Additions of Cu lead to an accelerated and even more enhanced
and Cu-modified [38] AlMg alloys, supersaturated vacancies are bound hardening effect in alloy 2, which is quite unusual for low temperatures.
to Mg atoms immediately after quenching. Elevated temperatures result The diffusivity of Cu atoms is limited at lower temperatures, so T-phase
in their gradual release and the formation of clusters which evolve into precursor precipitation is supposed to predominate the S-phase precip-
precipitates according to the predominant sequence. Due to their obvi- itation. As both phases were identified clearly after aging at 125 °C for
ous similarities, the alloys of this study are considered to behave in the 48 h, it is assumed that Cu somehow promotes the formation of T-
same way. phase precursors as suggested by Cao [29]. With the depletion of Mg
For the Cu-free alloy 1 the supposed sequence is SSSS → GP-zones due to the formation of T-phase precursors, the Cu/Mg ratio increases
(T") → T' → T-phase (Mg32(Al, Zn)49). This agrees well with the phase and S-phase formation is favored. Finally, both phases exist simulta-
identified after aging for 48 h at 125 °C, as no difference is made be- neously (Fig. 7c). A high density of precursor phase might also inhibit
tween T- and T'-phases due to their similar crystal structures [14,36]. the migration of solutes and free vacancies towards grain boundaries,
For the Zn- and Cu-containing alloy 2 both T-phase and S-phase were leading to narrower PFZs and finally resulting in higher hardness
observed in the same condition. Again, the distinction between S″, S′ [38,39].
and S-phase was not made in this study because of their similar struc-
tures [34,35]. 4.2. Two-step artificial aging
While at early aging times almost no hardening effect occurs at 175
°C for the Zn-only modified alloy 1 (Fig. 2a), a slight increase was mea- A key finding is the fact that applying a pre-aging treatment shifts
sured for alloy 2 with Cu additions (Fig. 2b). This agrees well with re- the hardening response of the investigated alloys towards an industri-
sults found by Cao [27] for aging at 180 °C. They conclude that this ally applicable time range.
effect might be the result of a favored Mg-Cu cluster formation, which During pre-aging at lower temperatures (100 °C) it is assumed that
occurs due to a high diffusivity of Cu and Mg atoms at this temperature. the formation of T-phase precursors is even more pronounced and oc-
This, however, leads to a relief in the lattice strain in the solid solution curs in higher number densities compared to single-step aging at 125

Fig. 5. Single-step artificial aging at 100 °C of alloy 1 (a) and alloy 2 (b) after water quenching (red line) and air cooling (blue dotted line) after solution heat treatment.
8 L. Stemper et al. / Materials and Design 181 (2019) 107927

°C. We link this to the increased Mg-supersaturation and the shorter dif- and g, the high number density of T-phase precursor formed during
fusion pathways. Cu additions seem to have an accelerating and pre-aging is obviously inherited by the following second hardening
boosting effect on the formation of T-phase precursors, similar to stage. For alloy 2 the two-step aging is assumed to proceed according
single-step aging of alloy 2 [28]. to the same pattern proposed by Cao [28,29], who investigated existing
Upon subsequent second-stage aging (Fig. 3a and b) previously phases and the distribution of alloying elements by TEM and atom
formed T-phase precursors may initially dissolve partly, causing a slight probe analysis in comparable alloys and heat treatment conditions. Ac-
decrease in hardness, but also act as nuclei for rapid hardening at ele- cording to his experimental results it was stated that Cu atoms are in-
vated artificial aging temperatures. As can be deduced from Fig. 6e corporated in clusters formed in the first aging step and, therefore, are

Fig. 6. HAADF microstructure of alloy 1 (left column) and alloy 2 (right column) after single-step artificial aging (125 °C/48 h, a, b, e, f) and two-step artificial aging (100 °C/3 h + 175 °C/3 h,
c, d, g, h), PFZ (red arrows).
L. Stemper et al. / Materials and Design 181 (2019) 107927 9

no longer available for the formation of S phase precursors. Thus, we alloys. While at the early aging stage previously formed but unstable
conclude that alloy 2 exhibits a very high number density of solely Cu- clusters may partially dissolve and cause hardness to decrease, continu-
containing T-phase in the peak aged condition (Fig. 7d). ing aging is slightly accelerated and hardening is intensified. This effect
At higher second-step temperatures the hardening response is fur- is thought to be caused by the growth of non-dissolved and, therefore,
ther accelerated, which is attributed to higher diffusion rates of solutes stable T-phase precursors, as suggested by Hou [30] for artificial pre-
in the matrix but is also accompanied by decreased peak hardness and aging. Interestingly, in contrast to their observations two weeks of nat-
early over-aging resulting from a loss of coherence and an undesired ural aging were able to generate these stable precipitates for the alloys
growth of the hardening phase. PFZ have been found to be less distinct investigated in this study; Hou found this effect to occur after pre-
in the two-step aged condition. This is attributed to the lower diffusion aging for 24 h at 90 °C, which may be explained by the different alloy
rates and higher number density of T-phase precursors caused by lower content. The reason for the extraordinary high strength level after pre-
temperature pre-aging. aging at room temperature might be the higher number density of clus-
ters, and finally precipitates, which evolve from these. A similar effect
4.3. Natural aging can also be found in lean 6xxx-series alloys, where clusters of Si and
Mg solute atoms, which formed during natural aging, are dissolving
The observed hardness increase at room temperature storage only slowly at subsequent artificial aging and/or act as nucleation sites
(Fig. 4a and b) is assumed to be related to a reduction of supersaturation for further precipitation. [42,43]
caused by the formation of clusters, as commonly found in 2xxx-, 6xxx-
and 7xxx-series alloys [40,41]. The effect can be desired or in some cases 4.4. Quench sensitivity
also undesired, due to instable mechanical properties in further process-
ing steps. Pre-aging treatment (100 °C for 3 h) shows a retarding effect, As expected, and typical of 7xxx- and 6xxx-series alloys, the age
which is more pronounced for the Cu-containing alloy 2. This might be hardening response is constrained for slow quenching in air compared
explained as follows. Due to the low temperature and resulting short to fast quenching in water [44–46]. The diminished supersaturation of
diffusion paths, cluster formation and growth are limited to narrow solutes and vacancies generates a lessened driving force for precipita-
spaces, resulting in a high number density of clusters. Likewise, super- tion of the hardening phase and makes a certain amount of solute un-
saturation decreases only slowly, so hardening can progress continu- available for hardening due to undesired coarse pre-precipitation
ously over time. The effect of pre-aging on natural aging can be linked while quenching.
to the above-mentioned formation of T-phase precursors. It decreases
the supersaturation of cluster-forming elements and thus stabilizes 5. Conclusions
the microstructure. While only minor differences are found in the natu-
ral aging behavior of alloy 1 and alloy 2, the effect of pre-aging on room This study investigated the aging behavior of two precipitation
temperature stability seems to be stronger for alloy 2. This can be ex- hardenable AlMg alloys processed via high pressure die casting. The
plained by the accelerating effect of Cu on the formation of the T- main conclusions for the AlMgZn and a Cu-added variant are:
phase precursors, similar to single-step aging.
Subsequent artificial aging (Fig. 4c and d) after two weeks of natural • Single-step artificial age hardening is relatively slow and is most ben-
aging achieves the highest measured hardening response for both eficial at lower aging temperature around 125 °C. The peak hardness is

Fig. 7. Simulated [11] and experimental diffraction patterns with beam direction b001N, Simulated diffraction patterns of S′/S″ (S)-phase (a), alloy 1 after single step aging at 125 °C for 48 h
(b), alloy 2 after single-step artificial aging at 125 °C/48 h (c) and after two-step artificial aging (100 °C/3 h + 175 °C/3 h) (d).
10 L. Stemper et al. / Materials and Design 181 (2019) 107927

related to T-phase precursor precipitates and is extended to a higher [4] F. Ostermann, Anwendungstechnologie Aluminium, Springer Berlin Heidelberg, Ber-
lin/Heidelberg, 2014https://doi.org/10.1007/978-3-662-05788-9.
level with the additions of Cu, resulting in a second hardening phase [5] S. Ji, D. Watson, Z. Fan, M. White, Development of a super ductile diecast Al–Mg–Si
(S-phase precursors). alloy, Mater. Sci. Eng. A 556 (2012) 824–833, https://doi.org/10.1016/j.msea.2012.
• Two-step artificial aging can shift the hardening peak into an industri- 07.074.
[6] M. Król, T. Tański, P. Snopiński, B. Tomiczek, Structure and properties of aluminium–
ally relevant time range. Longer pre-aging time increases the harden- magnesium casting alloys after heat treatment, J. Therm. Anal. Calorim. 127 (2017)
ing effect in the second stage. While a lower second-step temperature 299–308, https://doi.org/10.1007/s10973-016-5845-4.
increases the hardening response at longer times, a higher tempera- [7] B.-H. Lee, S.-H. Kim, J.-H. Park, H.-W. Kim, J.-C. Lee, Role of Mg in simultaneously im-
proving the strength and ductility of Al–Mg alloys, Mater. Sci. Eng. A 657 (2016)
ture results in a strong acceleration of precipitation but slightly less 115–122, https://doi.org/10.1016/j.msea.2016.01.089.
distinct strengthening. In peak aged condition, hardening results [8] M. Hino, S. Koga, S. Oie, M. Yanagawa, Properties of Al-Mg based alloys for automo-
from T-phase precursors in both alloys, but it is suggested that T- bile body panel, Kobelco Technol. Rev. 11 (1991) 1–5.
[9] O. Engler, C.D. Marioara, T. Hentschel, H.-J. Brinkman, Influence of copper additions
phase precursors incorporate certain amounts of Cu atoms for the
on materials properties and corrosion behaviour of Al–Mg alloy sheet, J. Alloys
Cu-containing alloy, leading to increased strength. Compd. 710 (2017) 650–662, https://doi.org/10.1016/j.jallcom.2017.03.298.
• Natural aging is pronounced and can be retarded by pre-aging, espe- [10] A. Alil, M. Popović, T. Radetić, M. Zrilić, E. Romhanji, Influence of annealing temper-
cially for the Cu-containing alloy. Interestingly, precursors formed ature on the baking response and corrosion properties of an Al–4.6wt% Mg alloy
with 0.54wt% Cu, J. Alloys Compd. 625 (2015) 76–84, https://doi.org/10.1016/j.
upon natural aging can even generate an extended peak hardening ef- jallcom.2014.11.063.
fect during subsequent artificial aging, compared to the artificial pre- [11] P. Ratchev, B. Verlinden, P. de Smet, P. van Houtte, Precipitation hardening of an Al–
aging treatments investigated. 4.2wt% Mg–0.6wt% Cu alloy, Acta Mater. (1998) 3523–3533.
[12] S.C. Wang, M.J. Starink, N. Gao, Precipitation hardening in Al–Cu–Mg alloys revisited,
The alloys investigated exhibit high potential for future deployment Scr. Mater. 54 (2006) 287–291, https://doi.org/10.1016/j.scriptamat.2005.09.010.
[13] P. Ratchev, B. Verlinden, P. de Smet, P. van Houtte, Artificial ageing of Al–Mg–Cu al-
in light-weight transportation applications. loys, Mater. Trans. (1999) 34–41.
[14] A. Bigot, P. Auger, S. Chambreland, D. Blavette, A. Reeves, Atomic scale imaging and
CRediT authorship contribution statement analysis of T' precipitates in Al-Mg-Zn alloys, Microsc. Microanal. Microstruct. 8
(1997) 103–113, https://doi.org/10.1051/mmm:1997109.
[15] X.B. Yang, J.H. Chen, J.Z. Liu, F. Qin, J. Xie, C.L. Wu, A high-strength AlZnMg alloy
Lukas Stemper: Conceptualization, Methodology, Investigation, Visuali- hardened by the T-phase precipitates, J. Alloys Compd. 610 (2014) 69–73, https://
zation, Writing - original draft. Bernhard Mitas: Investigation, Validation. doi.org/10.1016/j.jallcom.2014.04.185.
[16] C. Meng, Di Zhang, H. Cui, L. Zhuang, J. Zhang, Mechanical properties, intergranular
Thomas Kremmer: Investigation, Validation, Writing - original draft.
corrosion behavior and microstructure of Zn modified Al–Mg alloys, J. Alloys
Steffen Otterbach: Supervision. Peter J. Uggowitzer: Conceptualization, Compd. 617 (2014) 925–932, https://doi.org/10.1016/j.jallcom.2014.08.099.
Supervision, Writing - review & editing. Stefan Pogatscher: Project ad- [17] C.-Y. Meng, Zhang Di, P.-P. Liu, L.-Z. Zhuang, J.-s. Zhang, Microstructure characteriza-
tion in a sensitized Al–Mg–Mn–Zn alloy, Rare Metals 37 (2015) 129–135, https://
ministration, Conceptualization, Supervision, Writing - review & editing.
doi.org/10.1007/s12598-015-0665-4.
[18] J.-w. Zhao, B.-h. Luo, K.-j. He, Z.-h. Bai, B. Li, W. Chen, Effects of minor Zn content on
Declaration of Competing Interest microstructure and corrosion properties of Al−Mg alloy, J. Cent. South Univ. 23
(2016) 3051–3059, https://doi.org/10.1007/s11771-016-3368-6.
[19] M.C. Carroll, P.I. Gouma, M.J. Mills, G.S. Daehn, B.R. Dunbar, Effects of Zn additions on
None. the grain boundary precipitation and corrosion of Al-5083, Scr. Mater. 42 (2000)
335–340, https://doi.org/10.1016/S1359-6462(99)00349-8.
Acknowledgements [20] K. Matsumoto, Y. Aruga, H. Tsuneishi, H. Iwai, M. Mizuno, H. Araki, Effects of Zn ad-
dition and aging condition on serrated flow in Al-Mg alloys, MSF 794-796 (2014)
483–488, https://doi.org/10.4028/www.scientific.net/MSF.794-796.483.
Thankfully, this project was funded by the project “SMiLE” initiated [21] J. Yun, S. Kang, S. Lee, D. Bae, Development of heat-treatable Al-5Mg alloy sheets
by the German Federal Ministry of Education and Research and Audi with the addition of Zn, Mater. Sci. Eng. A 744 (2019) 21–27, https://doi.org/10.
1016/j.msea.2018.11.145.
AG. Financial support by the Christian Doppler Research Association,
[22] W.A. Curtin, D.L. Olmsted, L.G. Hector, A predictive mechanism for dynamic strain
the Austrian Federal Ministry for Digital and Economic Affairs and the ageing in aluminium-magnesium alloys, Nat. Mater. 5 (2006) 875–880, https://
National Foundation for Research, Technology and Development as well doi.org/10.1038/nmat1765.
[23] K. Matsumoto, Y. Aruga, H. Tsuneishi, H. Iwai, M. Mizuno, H. Araki, Effects of precip-
as AMAG rolling GmbH is gratefully acknowledged. We also thank
itation state on serrated flow in Al-Mg(-Zn) alloys, Mater. Trans. 57 (2016)
Manfred Brabetz for his support in conducting the TEM investigations. 1101–1108, https://doi.org/10.2320/matertrans.L-M2016814.
[24] M.C. Carroll, P.I. Gouma, G.S. Daehn, M.J. Mills, Effects of minor Cu additions on a Zn-
Data availability modified Al-5083 alloy, Mater. Sci. Eng. A 319-321 (2001) 425–428, https://doi.org/
10.1016/S0921-5093(00)02021-9.
[25] M.C. Carroll, R.G. Buchheit, G.S. Daehn, M.J. Mills, Optimum trace copper levels for
The raw/processed data required to reproduce these findings cannot SCC resistance in a Zn-modified Al-5083 alloy, MSF 396-402 (2002) 1443–1448,
be shared at this time as the data also forms part of an ongoing study. https://doi.org/10.4028/www.scientific.net/MSF.396-402.1443.
[26] C. Meng, Di Zhang, L. Zhuang, J. Zhang, Correlations between stress corrosion crack-
ing, grain boundary precipitates and Zn content of Al–Mg–Zn alloys, J. Alloys
CRediT authorship contribution statement Compd. 655 (2016) 178–187, https://doi.org/10.1016/j.jallcom.2015.09.159.
[27] C. Cao, Di Zhang, Z. He, L. Zhuang, J. Zhang, Enhanced and accelerated age hardening
response of Al-5.2Mg-0.45Cu (wt%) alloy with Zn addition, Mater. Sci. Eng. A 666
Lukas Stemper: Conceptualization, Methodology, Investigation, Vi- (2016) 34–42, https://doi.org/10.1016/j.msea.2016.04.022.
sualization, Writing - original draft. Bernhard Mitas: Investigation, Val- [28] C. Cao, Di Zhang, L. Zhuang, J. Zhang, Improved age-hardening response and altered
idation. Thomas Kremmer: Investigation, Validation, Writing - original precipitation behavior of Al-5.2Mg-0.45Cu-2.0Zn (wt%) alloy with pre-aging treat-
ment, J. Alloys Compd. 691 (2017) 40–43, https://doi.org/10.1016/j.jallcom.2016.
draft. Steffen Otterbach: Supervision. Peter J. Uggowitzer: Conceptual-
08.206.
ization, Supervision, Writing - review & editing. Stefan Pogatscher: [29] C. Cao, Di Zhang, X. Wang, Q. Ma, L. Zhuang, J. Zhang, Effects of Cu addition on the
Project administration, Conceptualization, Supervision, Writing - re- precipitation hardening response and intergranular corrosion of Al-5.2Mg-2.0Zn
(wt.%) alloy, Mater. Charact. 122 (2016) 177–182, https://doi.org/10.1016/j.
view & editing.
matchar.2016.11.004.
[30] S. Hou, P. Liu, Di Zhang, J. Zhang, L. Zhuang, Precipitation hardening behavior
References and microstructure evolution of Al–5.1 Mg–0.15Cu alloy with 3.0Zn (wt%)
addition, J. Mater. Sci. 53 (2018) 3846–3861, https://doi.org/10.1007/
[1] J. Hirsch, Recent development in aluminium for automotive applications, Trans. s10853-017-1811-1.
Nonferrous Metals Soc. China 24 (2014) 1995–2002, https://doi.org/10.1016/ [31] Pandat, Version 2018.1, www.computherm.com (database: PandatAl2018_TH).
S1003-6326(14)63305-7. [32] P. Villars, K. Cenzual, Mg32(Al,Zn)49 (Mg32Zn31.9Al17.1) crystal structure:
[2] J. Hirsch, Aluminium in innovative light-weight car design, Mater. Trans. 52 (2011) Datasheet from PAULING FILE multinaries edition, Springer Materials, Springer-
818–824, https://doi.org/10.2320/matertrans.L-MZ201132. Verlag Berlin Heidelberg & Material Phases Data System (MPDS), Switzerland,
[3] J. Hirsch, Automotive trends in aluminium - the European perspective, Mater. Forum 2012 , & National Institute for Materials Science (NIMS), Japan. https://materials.
28 (2004) 15. springer.com/isp/crystallographic/docs/sd_1252007.
L. Stemper et al. / Materials and Design 181 (2019) 107927 11

[33] G. Bergman, J.L.T. Waugh, L. Pauling, The crystal structure of the metallic phase [40] J.A. Österreicher, G. Kirov, S.S.A. Gerstl, E. Mukeli, F. Grabner, M. Kumar, Stabilization
Mg32(Al, Zn)49, Acta Cryst 10 (1957) 254–259, https://doi.org/10.1107/ of 7xxx aluminium alloys, J. Alloys Compd. 740 (2018) 167–173, https://doi.org/10.
S0365110X57000808. 1016/j.jallcom.2018.01.003.
[34] A. Charai, T. Walther, C. Alfonso, A.-M. Zahra, C.Y. Zahra, Coexistence of clusters, GPB [41] S. Jin, T. Ngai, L. Li, Y. Lai, Z. Chen, A. Wang, Influence of natural aging and pre-
zones, S″-, S′- and S-phases in an Al–0.9% Cu–1.4% Mg alloy, Acta Mater. 48 (2000) treatment on the precipitation and age-hardening behavior of Al-1.0Mg-0.65Si-
2751–2764, https://doi.org/10.1016/S1359-6454(99)00422-X. 0.24Cu alloy, J. Alloys Compd. 742 (2018) 852–859, https://doi.org/10.1016/j.
[35] L. Kovarik, M.K. Miller, S.A. Court, M.J. Mills, Origin of the modified orientation rela- jallcom.2017.10.005.
tionship for S(S″)-phase in Al–Mg–Cu alloys, Acta Mater. 54 (2006) 1731–1740, [42] C.S.T. Chang, I. Wieler, N. Wanderka, J. Banhart, Positive effect of natural pre-ageing
https://doi.org/10.1016/j.actamat.2005.11.045. on precipitation hardening in Al-0.44at% Mg-0.38at% Si alloy, Ultramicroscopy 109
[36] H. Löffler, I. Kovcs, J. Lendvai, Decomposition processes in Al-Zn-Mg alloys, J. Mater. (2009) 585–592, https://doi.org/10.1016/j.ultramic.2008.12.002.
Sci. 18 (1983) 2215–2240, https://doi.org/10.1007/BF00541825. [43] S. Pogatscher, H. Antrekowitsch, T. Ebner, P. Uggowitzer, The role of co-clusters in
[37] A.J. Perry, Solute-vacancy interaction energies and the effect of 0.009 at.% Mg on the the artificial aging of AA6061 and AA6060, Light Metals 2012, pp. 415–420.
ageing kinetics of an Al-4.01 at.% Zn alloy, Acta Metall. 14 (1966) 1143–1156, [44] M. Conserva, P. Fiorini, Interpretation of quench-sensitivity in Al-Zn-Mg-Cu alloys,
https://doi.org/10.1016/0001-6160(66)90232-X. MT 4 (1973) 857–862, https://doi.org/10.1007/BF02643097.
[38] R.K.W. Marceau, G. Sha, R. Ferragut, A. Dupasquier, S.P. Ringer, Solute clustering in [45] M.J. Starink, B. Milkereit, Y. Zhang, P.A. Rometsch, Predicting the quench sensitivity
Al–Cu–Mg alloys during the early stages of elevated temperature ageing, Acta of Al–Zn–Mg–Cu alloys, Mater. Des. 88 (2015) 958–971, https://doi.org/10.1016/j.
Mater. 58 (2010) 4923–4939, https://doi.org/10.1016/j.actamat.2010.05.020. matdes.2015.09.058.
[39] T. Ogura, S. Hirosawa, T. Sato, Quantitative characterization of precipitate free zones [46] B. Milkereit, M.J. Starink, Quench sensitivity of Al–Mg–Si alloys, Mater. Des. 76
in Al–Zn–Mg(–Ag) alloys by microchemical analysis and nanoindentation measure- (2015) 117–129, https://doi.org/10.1016/j.matdes.2015.03.055.
ment, Sci. Technol. Adv. Mater. 5 (2004) 491–496, https://doi.org/10.1016/j.stam.
2004.02.007.

You might also like