You are on page 1of 10

Journal of Alloys and Compounds 530 (2012) 71–80

Contents lists available at SciVerse ScienceDirect

Journal of Alloys and Compounds


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

Effects of Sc and Zr microalloying additions on the microstructure and


mechanical properties of new Al–Zn–Mg alloys
Ying Deng a,∗ , Zhimin Yin a , Kai Zhao a , Jiaqi Duan a , Zhenbo He b
a
School of Materials Science and Engineering, Central South University, Hunan, Changsha 410083, China
b
Northeast Light Alloy Co. Ltd., Hei Longjiang, Harbin 150060, China

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

Article history: Effects of Sc and Zr microalloying additions on the microstructure and mechanical properties of new
Received 23 December 2011 Al–Zn–Mg alloys alloyed with a small amount of copper were investigated comparatively by tensile
Received in revised form 26 March 2012 tests and microscopy methods. Compared with the strength of peak-aged Al–Zn–Mg alloy, the yield
Accepted 26 March 2012
strength increased by 66 MPa after adding 0.10 wt.% Sc and 0.10 wt.% Zr, and improved by 96 MPa after
Available online 5 April 2012
adding 0.25 wt.% Sc and 0.10 wt.% Zr, respectively. Introduction of 0.10 wt.% Zr to Al–Zn–Mg alloy, the
grain refinement effect of scandium did not occur by adding only 0.10 wt.% Sc, but appeared with an
Keywords:
increase in scandium content to 0.25 wt.%. In the presence of 0.10 wt.% Zr, antirecrystallized effect of
Aluminum alloys
Precipitation
scandium appeared at 0.10 wt.% scandium concentration, and enhanced with an increase in the content
Mechanical properties of scandium. Main aging precipitates were two kinds of GP zones in under-aged alloys, and ␩ phases in
Microstructure peak-aged alloys, respectively. Small additions of Sc and Zr did not retard or suppress the formation of
Electron microscopy aging precipitates in Al–Zn–Mg–Sc–Zr alloys. The strengthening effect from aging precipitates was much
stronger than from Sc and Zr microalloying additions in Al–Zn–Mg–Sc–Zr alloys. In addition, the most
important strengthening mechanisms of Sc and Zr microalloying in aged Al–Zn–Mg alloys was Orowan
strengthening of secondary Al3 (Sc, Zr) particles.
© 2012 Elsevier B.V. All rights reserved.

1. Introduction that was attributed to precipitation of Zr-enriched outer shells onto


the Al3 Sc precipitates, leading to a peak microhardness of 618 MPa
Of all microalloying additions to Al, Sc and Zr offer the great- at 400 ◦ C for Al–0.06Sc–0.06Zr. Van Dalen [14] and Booth-Morrison
est potential for developing new lightweight structural materials [15] indicated that additions of Zr and Sc improved coarsening
with excellent mechanical properties, good welding performance resistance of Al3 (Sc, Zr) (L12 ) precipitates compared with Al3 Sc
and desirable corrosion and creep resistance [1–6]. Mukhopad- (L12 ).
hyay [7] found that in Al–Zn–Mg–Cu alloy containing both Zr High-strength thermally strengthened weldable alloys based on
and Sc as the dispersoid forming elements, the recrystallized the Al–Zn–Mg system are used extensively for structural applica-
grains were very stable against coarsening, and the total elon- tions in aerospace industry [16–19], and recently reported to be
gation value can reach 916%, exhibiting high superplasticity. alloyed with a small amount of copper, which improves consider-
Dev [8] observed that addition of scandium led to very signifi- ably the resistance of the base metal and especially that of welded
cant grain refinement in the fusion zone especially for scandium joints to corrosion cracking and slow disruption. Though the pres-
levels greater than the eutectic composition (0.55 wt.%). Desir- ence of copper intensifies the formation of hot cracks in welding of
able service performance from additions of minor Sc and Zr in aluminum alloys, this drawback is eliminated by the introduction
aluminum alloys is mainly attributed to the presence of coher- of a combined scandium and zirconium additive [20].
ent, L12 -ordered Al3 (Sc, Zr) particles [4,9,10]. The reports about However, the systematical and comparative study on the
the aluminum alloys with scandium and zirconium additive synergistic effect of Sc and Zr microalloying additions on the
s are available but mainly concentrated on the precipitation behav- microstructure and properties of new Al–Zn–Mg alloys alloyed
iors of Al3 (Zr, Sc) particles [11,12]. Knipling [13] observed that, with a small amount of copper is not available. The purpose of this
above 325 ◦ C, Zr additions provided a secondary strength increase work is to study the relationship between the mechanical behavior
and microstructure characteristics of new Al–Zn–Mg alloys with
different scandium and zirconium additives, to understand the
∗ Corresponding author. Tel.: +86 731 88830262; fax: +86 731 88830262. microalloying mechanisms of Sc and Zr additions in new Al–Zn–Mg
E-mail address: csudengying@163.com (Y. Deng). alloys.

0925-8388/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jallcom.2012.03.108
72 Y. Deng et al. / Journal of Alloys and Compounds 530 (2012) 71–80

Table 1
Chemical composition of studied alloys (in wt.%).

Alloy Zn Mg Sc Zr Cu Mn Si Fe Al

Al–Zn–Mg 5.39 1.91 – – 0.34 0.33 0.08 0.16 Bal.


Al–Zn–Mg–0.10Sc–0.10Zr 5.36 1.90 0.10 0.10 0.25 0.32 0.09 0.19 Bal.
Al–Zn–Mg–0.25Sc–0.10Zr 5.41 1.98 0.25 0.10 0.33 0.32 0.11 0.18 Bal.

2. Experimental procedures Zr, respectively, and improved by 96 MPa and 58 MPa by adding
0.25 wt.% Sc and 0.10 wt.% Zr, respectively, meanwhile the elon-
Three kinds of Al–Zn–Mg alloys with and without Sc and Zr additions were
gations all kept above 12% in two peak-aged Al–Zn–Mg–Sc–Zr
applied for comparative research. Semi-continuous ingots, with 172 mm in diam-
eter, were provided by Northeast Light Alloy Co. Ltd. Table 1 shows their chemical alloys. The strength increments can be explained by strength-
compositions. After homogenization treatment (350 ◦ C × 8 h + 470 ◦ C × 12 h: the ening due to additions of Sc and Zr. Moreover, with the
purpose of the pretreatment at 350 ◦ C for 8 h is to promote the nucleation of fine increase of scandium additions, the strength increased and the
and disperse Al3 (Zr, Sc) particles [12,13,21]; the aim of the final treatment at 470 ◦ C elongation decreased. The yield strength increment caused by
for 12 h is to eliminate non-equilibrium eutectic phases), the ingots with dimension
of Ø172 mm × 70 mm, were inter-annealed at 420 ◦ C for 4 h, then hot rolled to 7 mm
aging treatment was up to 256 MPa, 253 MPa and 260 MPa in
thick plates immediately. Then the hot rolled plates were annealed at 420 ◦ C for 1 h, Al–Zn–Mg alloy, Al–Zn–Mg–0.10 wt.% Sc–0.10 wt.% Zr alloy and
and cold rolled to 2 mm sheets. The sheets were subjected to solution treatment Al–Zn–Mg–0.25 wt.% Sc–0.10 wt.% Zr, respectively. The substantial
at 470 ◦ C for 1 h, followed by water quenching, and then aged at 120 ◦ C for 24 h. increase in the yield strength after aging was evidently related to
The selection of heat treatment parameters was based on the results reported in
age strengthening. Besides, obviously, strengthening effect from
reference [2].
In order to investigate the effects of scandium and zirconium microalloying addi- aging precipitates (253 MPa and 260 MPa) was much larger than
tions in Al–Zn–Mg alloys, the microstructure and properties of Al–Zn–Mg alloys with from microalloying additions of Sc and Zr (66 MPa and 96 MPa) in
different Sc and Zr contents were studied comparatively by hardness measurements, Al–Zn–Mg alloy.
tensile tests, scanning electron microscopy (SEM), energy dispersive spectroscopy
(EDS), transmission electron microscopy (TEM), electron back scattered diffraction
(EBSD) technique and high resolution electron microscopy (HREM) methods. Hard- 3.2. Effect of Sc and Zr microalloying additions on grain
ness measurements were finished on an HW187.5 Brinell’s machine. The mechanical
microstructures
testing utilized transverse orientation (normal to rolling direction) specimens. Ten-
sile properties were performed on a CSS-44100 electronic universal testing machine
with 2 mm/min loading speed. Yield strength of materials was identified at 0.2% Fig. 1 is the orientation map with boundaries for homoge-
plastic strain. A minimum of six hardness measurements and mechanical tests was nized and aged alloys. The fractions of grain boundaries with
performed on each sample, and their statistical scatter determined the measure-
different misorientation angles are shown in Fig. 2. The orienta-
ment errors. All of the experimental errors reported here represented one standard
deviation from the mean. SEM observations were carried out on a Quanta MK2-200 tion maps of three homogenized alloys consisted of the grains
SEM, operating at 20 kV. Thin foils for transmission electron microscope (TEM) and with average size of 78.3 ␮m, 79.6 ␮m and 58.4 ␮m in Al–Zn–Mg
high resolution electron microscope (HRTEM) observations were sectioned from alloy, Al–Zn–Mg–0.10Sc–0.10Zr alloy, Al–Zn–Mg–0.25Sc–0.10Zr
the alloys under different conditions. The foils were prepared by twin-jet electro- alloy respectively. This indicated that making additions of 0.25 wt.%
polishing at 20 V in a solution of 30% nitric acid and 70% methanol solution cooled
Sc and 0.10 wt.% Zr to Al–Zn–Mg alloys can refine the grains
to −30 ◦ C and observed on a TECNAIG2 20 electron microscope and a JEM-3010
electron microscope respectively, both with an acceleration voltage of 200 kV. EBSD of ingots, while the modifying effect did not occur when
analyses were performed using a Sirion 200 field emission gun scanning electron the content of Sc was only 0.10 wt.%. The grain boundaries
microscope equipped with EBSD system. EBSD data were subsequently analyzed by were all mostly characterized by HAGB. After solution-aging
the OIM Analysis 5 software. The orientation uncertainty or “orientation noise” in
treatment, recrystallization occurred in Al–Zn–Mg alloy, where
deformed samples (aged alloy) was greater than in homogenized samples. Therefore
in order to avoid spurious boundaries, misorientations below 2◦ were not measured. grain boundaries were mostly characterized by HAGB. However,
This limit was used for all samples in order to provide consistent quantitative data. Al–Zn–Mg–0.10Sc–0.10Zr and Al–Zn–Mg–0.25Sc–0.10Zr peak-
Boundaries with misorientations between 2◦ and 15◦ were defined as low angle grain aged alloys still remained unrecrystallized, fiber-like structures.
boundaries (LAGB) and those of misorientation >15◦ as high angle grain boundaries The fiber-like structures were made up of sub-grains whose
(HAGB).
boundaries were characterized by LAGB. Moreover, an increase in
scandium content promoted an increase in LAGB, exhibiting the
3. Results misorientation angles between 2◦ and 15◦ , associated to substruc-
ture (Figs. 1 and 2). The grain sizes were 8.26 ␮m, 1.36 ␮m and
3.1. Effect of Sc and Zr microalloying additions on tensile 0.93 ␮m in aged Al–Zn–Mg alloy, Al–Zn–Mg–0.10Sc–0.10Zr alloy
properties and Al–Zn–Mg–0.25Sc–0.10Zr alloy, respectively.
SEM images in Fig. 3 show microstructures of as-cast and
Table 2 shows tensile properties of the Al–Zn–Mg alloys with homogenized Al–Zn–Mg and Al–Zn–Mg–0.25Sc–0.10Zr alloys.
different Sc and Zr additions under different conditions. Com- Serious segregation existed in the two studied as-cast alloys. EDS
pared with the mechanical properties of peak-aged Al–Zn–Mg results are shown in Table 3. According to EDS analyses, the white
alloy, the yield strength and the ultimate tensile strength increased secondary phases as shown in Fig. 3(a)A and Fig. 3(b)C, were both
by 66 MPa and 31 MPa after adding 0.10 wt.% Sc and 0.10 wt.% enriched in Zn and Mg elements non-equilibrium phases. The grey

Table 2
Mechanical properties of Al–Zn–Mg alloys with different Sc and Zr additives under different processing conditions.

Alloys Condition YS (MPa) UTS (MPa) EI (%)

Al–Zn–Mg Solution treated 183 ± 3 328 ± 1 25.2 ± 0.1


Peak aged 439 ± 2 498 ± 1 16.5 ± 0.7

Al–Zn–Mg–0.1Sc–0.1Zr Solution treated 252 ± 2 378 ± 3 23.6 ± 0.4


Peak aged 505 ± 3 529 ± 2 14.1 ± 0.6

Al–Zn–Mg–0.25Sc–0.1Zr Solution treated 275 ± 1 397 ± 3 23.3 ± 0.3


Peak aged 535 ± 5 556 ± 2 12.1 ± 0.4
Y. Deng et al. / Journal of Alloys and Compounds 530 (2012) 71–80 73

Fig. 1. Orientation maps of specimens under different conditions. Al–Zn–Mg alloy: (a) homogenized and (d) aged; Al–Zn–Mg–0.10Sc–0.10Zr alloy: (b) homogenized and (e)
aged; Al–Zn–Mg–0.25Sc–0.10Zr alloy: (c) homogenized and (f) aged; (g) representation of different color lines used to identify the boundaries with different misorientation
angles in (a)–(f); (h) direction of specimens in (d)–(f). (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this
article.)

100
Boundary levels (misorientation angle)
phases as shown in Fig. 3(a)B and Fig. 3(b)D were enriched in Fe
LAB 2-15° and Mn elements indissoluble impurity phases. After homogeniza-
HAB 15-180° tion treatment, enriched in Zn and Mg non-equilibrium phases
80 dissolved into matrix, and only small indissoluble impurity phases
remained (Fig. 3(c)F and (d)G). Compared with Al–Zn–Mg alloy, an
additional typical characteristic of Al–Zn–Mg–0.25Sc–0.10Zr alloy
60 was the presence of primary Al3 (Sc, Zr) particles, distributed ran-
Frequency/%

domly in the center of some grains (Fig. 3(b)E and (d)H).


TEM images in Fig. 4 show microstructures of the three
40 aged alloys. Compared with the microstructure of Al–Zn–Mg
alloy consisted of recrystallized grains, the structure of
Al–Zn–Mg–0.10Sc–0.10Zr and Al–Zn–Mg–0.25Sc–0.10Zr aged
20 alloys both consisted of micro-scaled sub-grain structures. More-
over, with an increase in the content of scandium, the size of
sub-grains decreased. The results are in accordance with EBSD
0 results.
Al-Zn-Mg 0.10 wt%Sc 0.25 wt%Sc Al-Zn-Mg 0.10 wt%Sc 0.25 wt%Sc

Homogenized Aged
3.3. Effect of Sc and Zr microalloying additions on age hardening
Fig. 2. Frequency of different misorientation angles in the three alloys under differ-
ent conditions. Fig. 5 shows the effect of aging time on the hardness of the
three alloys. The three alloys all exhibited typical aging hardening

Table 3
Chemical composition of the secondary phases in Fig. 2 (in at.%).

Phase Al Zn Mg Cu Fe Si Mn Sc Zr Ti

A 65.69 13.94 16.72 3.76 0 0 0 0 0 0


B 78.77 1.64 2.01 0.84 8.86 3.96 3.93 0 0 0
C 68.20 13.07 15.65 3.09 0 0 0 0 0 0
D 83.04 1.42 1.92 0.63 7.49 3.61 1.90 0 0 0
E 76.46 1.83 0 0 0 0 0 11.61 8.73 1.38
F 80.72 1.10 1.10 1.05 10.22 3.25 2.56 0 0 0
G 83.14 1.18 1.35 0.73 8.82 3.27 1.51 0 0 0
H 75.52 1.74 0 0 0 0 0 11.85 9.17 1.73
74 Y. Deng et al. / Journal of Alloys and Compounds 530 (2012) 71–80

Fig. 3. SEM microstructures of the two studied alloys under different conditions. Al–Zn–Mg alloy: (a) as-cast and (c) homogenized; Al–Zn–Mg–0.25Sc–0.10Zr alloy: (b)
as-cast and (d) homogenized.

behaviors. Hardness vs. aging time curves in Fig. 5 exhibit three 3.4. Effect of Sc and Zr microalloying additions on aging
regions: (1) rapid increase in hardness (under-aging); (2) plateau in precipitates in Al–Zn–Mg alloy
hardness (peak-aging); and (3) decrease in hardness (over-aging).
Three alloys were all in peak-aged T6 state after 24 h of aging at TEM images in Fig. 6 show microstructures of the three aged
120 ◦ C. Seen from Fig. 5, the hardness increased by adding Sc and alloys. A large number of fine precipitates were formed homo-
Zr to Al–Zn–Mg alloys, and the hardness increment was uniform geneously in grains and along grain boundaries (Fig. 6(a)–(c))
over the entire ageing process. This indicated that hardening from in under aged alloys. With increasing aging time, short rod-like
minor Sc and Zr was independent of hardening caused by other particles gradually coarsened within grains, and grain boundary
aging precipitates in Al–Zn–Mg alloy. precipitates were formed in peak aged alloys (Fig. 6(d)–(f)). By

Fig. 4. Bright field TEM microstructures of the three aged alloys: (a) Al–Zn–Mg, (b) Al–Zn–Mg–0.10Sc–0.10Zr and (c) Al–Zn–Mg–0.25Sc–0.10Zr.
Y. Deng et al. / Journal of Alloys and Compounds 530 (2012) 71–80 75

was that the size of grain boundary precipitates and the width of
180 the precipitate free zone were smaller in Al–Zn–Mg–Sc–Zr alloys
(Fig. 6(g)–(i)).

160
3.5. Precipitates in Al–Zn–Mg–Sc–Zr alloys
Al-Zn- Mg
Hardness / HB

140 Al-Zn- Mg-0. 10Sc- 0.10Zr


TEM images in Fig. 7 show precipitates in Al–Zn–Mg–Sc–Zr
Al-Zn- Mg-0. 25Sc- 0.10Zr alloys under different conditions. With semi-continuous casting
and subsequent cooling of ingots, a considerable part of scan-
120 dium and zirconium was fixed in supersaturated solid solution.
During subsequent homogenization treatment, scandium and zir-
conium precipitated from supersaturated ␣(Al) matrix in the form
100 of secondary Al3 (Sc, Zr) particles [9,22–26]. The average size of
Al3 (Sc, Zr) precipitates was estimated from direct measurements
on TEM negatives, and was determined to be about 20 nm in
80
0 10 20 30 40 50 homogenized Al–Zn–Mg–Sc–Zr alloys (Fig. 7(a)–(b)). The presence
Aging time/ h
of Ashby–Brown contrast for bright-field images, as well as the
superstructure reflections like (1, 0, 0) and (1, 1, 0), confirmed
Fig. 5. Age hardening curves. that this particle was Al3 (Sc, Zr) particle with an Ll2 cubic crystal
structure, which was coherent with the Al matrix. The cold rolled
microstructures mainly consisted of dislocation substructures. In
comparison, it was notable that there was almost no difference of addition, many fine spherical particles without Ashby–Brown con-
the size and density of aging precipitates within grains between trast were observed in bright-field images. The [0 1 1] orientation
Al–Zn–Mg alloy and Al–Zn–Mg–Sc–Zr alloy (Fig. 6(a)–(i)). The only diffraction pattern from those particles were performed, show-
difference between the Al–Zn–Mg alloy and Al–Zn–Mg–Sc–Zr alloy ing characteristic super-lattice spots from coherent Al3 (Sc, Zr)

Fig. 6. Bright field microstructural evolution of the three alloys aged at 120 ◦ C. Al–Zn–Mg alloy: (a) under aged, (d) peak aged and (g) over aged; Al–Zn–Mg–0.10Sc–0.10Zr
alloy: (b) under aged, (e) peak aged and (h) over aged; Al–Zn–Mg-0.25Sc–0.10Zr alloy: (c) under aged, (f) peak aged and (i) over aged.
76 Y. Deng et al. / Journal of Alloys and Compounds 530 (2012) 71–80

Fig. 7. Precipitates in Al–Zn–Mg–Sc–Zr alloys under different conditions. Al–Zn–Mg–0.10Sc–0.10Zr alloy: (a) homogenized, (e) aged at 120 ◦ C for 1 h and (i) peak aged;
Al–Zn–Mg–0.25Sc–0.10Zr alloy: (b) homogenized, (c) hot-rolled, (d) solution-treated, (f) aged at 120 ◦ C for 1 h and (j) peak aged; (g and h) HREM images: taken along the
[1 0 0] and [1 1 0] zone axis, aged at 120 ◦ C for 1 h.

precipitates. The loss of Ashby–Brown contrast around secondary field images appeared again for the release of strain. Al3 (Sc, Zr)
Al3 (Sc, Zr) precipitates in rolled alloys was due to the interfer- particles strongly pinned sub-grain boundaries (Fig. 7(d)). The
ence of dislocation strain field (Fig. 7(c)). After solution treatment, bright field and HREM images of the under-aged alloy are shown
unrecrystallized micrometer-sized sub-grains remained. More- in Fig. 7(e) and (f) and (g) and (h) respectively. In bright field
over, Ashby–Brown contrast around Al3 (Sc, Zr) particles for bright image under the just Bragg condition, Al3 (Sc, Zr) particles were still
Y. Deng et al. / Journal of Alloys and Compounds 530 (2012) 71–80 77

Fig. 7. (Continued ).

coherent with matrix judged from Ashby and Brown contrast which projection. Therefore, ␩’ phases and Al3 (Sc, Zr) particles were
appeared around a spherical coherent particle in Fig. 7(e) and (f). the main strengthening phases in peak-aged Al–Zn–Mg–Sc–Zr
The size of coherent Al3 (Sc, Zr) particles kept unchanged during alloy.
heat treatment and processing in our research, which was sup-
ported by the work of Iwamura and Miura [27], who reported 4. Discussion
that accelerated Al3 (Sc, Zr) particle growth was observed only at
particle diameters larger than 80 nm, while delay in particle coars- 4.1. The modifying effect attained due to Sc and Zr microalloying
ening was noticed in the particle diameter range of 30–80 nm. additions to Al–Zn–Mg alloy
Besides, high resolution images were taken in the Al-projections
1 0 0 and 1 1 0, as shown in Fig. 7 (g) and (h). Contrast fea- In the former Soviet Union, additions of the transition element
tures in the [1 0 0]Al projection, Fig. 7 (g), were identified as scandium (Sc) have been used for a number of years as an alterna-
GP(I)-zones, which appeared as spherical precipitates, 2–8 nm tive grain refiner to Ti/TiB2 , particularly in the context of welding.
wide. In the [1 1 0]Al projection, Fig. 7(h), the thin objects were Grain refinement due to the additions of scandium has been pre-
identified as GP(II)-zones, 3–7 nm wide, which were assumed viously shown to arise from the precipitation of primary Al3 Sc
to be an early stage in the formation of ␩ -precipitates. The particles in the melt, which act as potent nucleation sites for alu-
GP(II)-zones were described as zinc-rich layers on {1 1 1}-planes, minum grains [30]. As the aluminum-rich end of the Al–Sc phase
with internal order in the form of elongated 1 1 0 domains diagram is dominated by a eutectic reaction between the liquid, ␣-
[28]. Therefore, the two types of GP-zones and Al3 (Sc, Zr) parti- Al, and Al3 Sc, the grain refinement mechanism only operates with
cles were the main strengthening precipitates in the under-aged scandium addition levels greater than the eutectic composition
Al–Zn–Mg–Sc–Zr alloy. In peak-aged alloy, besides a large number (0.55 wt.% Sc). Taking into account that the cost of the Al–Sc alloy-
of coherent secondary Al3 (Sc, Zr) particles, lots of fine precipi- ing combination is high, it is economically inexpedient to introduce
tates were distributed homogenously in matrix. The SAD pattern this high content of scandium into industrial alloys (even for air-
of peak aged alloys in [1 1 2]Al projection was shown in Fig. 7(j). craft and spacecraft). Fortunately, the minimum Sc level required
Besides the spots of Al3 (Sc, Zr) particles, there were some spots for refinement can be significantly reduced by the addition of Zr
from ␩ phase, which were identified by referring to the model in [20,31]. The previous reports [32,33] showed that when Zr addi-
some reported results [29]: (10.l) and (20.l) rows in the [1 1 2]Al tions were not made, the modifying effect of scandium appears at
78 Y. Deng et al. / Journal of Alloys and Compounds 530 (2012) 71–80

concentrations above 0.55 wt.% (at hypereutectic concentrations), mechanism should be responsible for the Al3 (Sc, Zr) particle-
while in the presence of zirconium (about 0.10 wt.%), the modifying induced
effect manifests itself at scandium concentration above 0.18 wt.%. strengthening in our studies (20 nm). Therefore, a increase in
In our research, in the case of 0.10 wt.% Zr additions into Al–Zn–Mg the YS after peak aging (66 MPa and 96 MPa) caused by adding
alloy, the grain refinement effect can be observed when the content minor Sc and Zr was evidently related to refining strengthening
of scandium was 0.25 wt.%, but did not appear at scandium con- from primary Al3 (Sc, Zr) particles formed during the solidification
centration of 0.10 wt.% Sc. This result was in good accordance with in Al–Zn–Mg–0.25Sc–0.10Zr alloy, substructure strengthening
the references [32,33]. In fact, an additive consisting of 0.25 wt.% and Orowan strengthening caused by the secondary Al3 (Sc, Zr)
Sc and 0.10 wt.% Zr replaces the introduction of 0.5–0.6 wt.% Sc particles precipitated during homogenization in Al–Zn–Mg–Sc–Zr
from the standpoint of modifying the cast grain structure of alu- alloys. Besides, compared with the mechanical properties reported
minum alloys. This saves expensive scandium. The high level of in reference [2], the alloy gained 11 MPa yield strength increment
grain refinement was caused by primary Al3 (Sc, Zr) particles, solid- by pretreatment at 350 ◦ C for 8 h during homogenization in aged
ified firstly during solidification, as observed in Fig. 3(b) and (d), Al–Zn–Mg–0.25Sc–0.10Zr alloy.
which were active centers for the generation of new grains of solid Wu [26] indicates that the formation of GP(II) zones is sup-
solution [34]. pressed in the presence of zirconium that leads to a reduction in the
amount of ␩ formed during subsequent ageing. However, accord-
4.2. The antirecrystallized effect attained due to making Sc and Zr ing to the references [28], GP(I)-zones are formed over a wide
microalloying additions to Al–Zn–Mg alloy temperature range, from room temperature to 140–150 ◦ C, inde-
pendently of quenching temperature. The GP(I)-zones are coherent
As practice of the work with commercial scandium enriched with the aluminum matrix, with internal ordering of Zn and Al/Mg
aluminum alloys shown, optimum scandium content, which on the matrix lattice, suggested to be based on AuCu(I)-type sub-
guarantees retention of a nonrecrystallized structure after heat unit, and anti-phase boundaries. GP(II) are formed after quenching
treatment of semiproducts, depends on an alloy system. Al–Zn–Mg from temperatures above 450 ◦ C, by aging at temperatures above
system based alloys are susceptible to recrystallization to a lesser 70 ◦ C. The solution-aging treatment in our studies is solution
extent, and scandium content within a range of 0.15–0.20 wt.% treated at 470 ◦ C for 1 h, followed by water quenching and then aged
ensures retention of a non-recrystallized structure of quenched at 120 ◦ C for 24 h, which satisfies the condition of forming two kinds
semiproducts [33]. In our studies, in the presence of 0.10 wt.% Zr, a of GP zones. Moreover, from our observation, two kinds of GP zones
nonrecrystallized structure remains at scandium concentration of both appear in our studied alloy during the early stage of aging. Fur-
0.10 wt.% in Al–Zn–Mg alloy. Therefore, in the presence of 0.10 wt.% thermore, observed from the microstructures of two aged alloys,
Zr, the scandium content ensuring retention of a non-recrystallized small additions of Sc and Zr do not affect the size and density of
structure is lower than Zr-free Al–Zn–Mg alloy. Moreover, antire- the aging precipitates within grains during aging treatment, except
crystallized effect of scandium enhances by increasing scandium the grain boundary precipitates and precipitate free zone. There-
additions. High antirecrystallized effectiveness of Sc and Zr addi- fore, we can infer that minor additions of Zr and Sc to Al–Zn–Mg
tions is attributed to high density of the fine, coherent, secondary alloy do not retard or suppress the formation of aging harden-
Al3 (Sc, Zr) dispersoid particles precipitated during decomposition ing precipitates. This inference is confirmation for the mechanical
of solid solution homogeneously, as shown in Fig. 7(a) and (b). property results that the YS increment from aging precipitation is
Coherent Al3 (Sc, Zr) particles in the diameter of about 20 nm have a almost the same among the three studied alloys (256 MPa, 253 MPa
good thermal stability and drastic antirecrystallized effect. As they and 260 MPa respectively). Besides, the strengthening effect from
strongly pin dislocation and grain boundaries during heat treat- aging precipitates (253 MPa and 260 MPa) is much stronger than
ment and processing, a nonrecrystallized structure (substructure) from Sc and Zr microalloying additions (66 MPa and 96 MPa) in
still remains in aged Al–Zn–Mg–Sc–Zr alloys (Figs. 1, 2 and 4). Al–Zn–Mg–Sc–Zr alloys.

4.3. Effects of Sc and Zr microalloying additions on mechanical 4.4. Strengthening mechanisms of Sc and Zr additions
properties and aging precipitation
According to the discussion in Sections 4.1–4.3, the strength-
Small additions of Sc and Zr enable a superior combination of ening mechanisms of Sc and Zr microalloying additions in aged
higher strength levels and acceptable ductility in aged alloy. Start- Al–Zn–Mg alloys can be mainly divided into Orowan strengthening
ing from refining the grain size of cast aluminum alloys, additions derived from secondary Al3 (Sc, Zr) particles and gain boundary or
of Sc and Zr also increase the resistance to recrystallization during subgrain boundary strengthening. However, the most important
hot working and introduce additional strengthening through the mechanism of Sc and Zr microalloying additions in our stud-
formation of fine coherent secondary Al3 (Sc, Zr) particles. Homo- ied alloys is unclear. Therefore, we need to estimate different
geneous distribution of these particles promotes formation of a strengthening contributions. Due to the limitation of experimental
stable refined subgrain structure during deformation processing conditions, we just provided coarse estimation here. We assumed
and heat treatment, which gives an additional increase in strength. that the thickness of the TEM-foils in the area of investigation was
A strength increase due to the secondary Al3 (Sc, Zr) precipitation about 80 nm, which was the estimation from TEM sample prepa-
can generally occur by one of two mechanisms, depending on ration. The sizes (diameters) and the number density of the Al3 (Sc,
the particle size. These are (i) shearing of small precipitates, Zr) particles were estimated from TEM images taken at magnifi-
which involves order strengthening, coherency strengthening and cations of 195 000–380 000. The microstructural features of Al3 (Sc,
modulus mismatch strengthening; and (ii) the Orowan bypass Zr) particles in studied aged alloys were shown in Table 4.
mechanism for larger particles, when dislocations cannot move
through but loop around the particles [35]. With an increase in par- 4.4.1. Orowan strengthening of Al3 (Sc, Zr) particles
ticle diameter, the YS increases when the first mechanism operates, The increase in the YS due to the Orowan strengthening,  Or ,
while it decreases when the Orowan mechanism operates. Transi- was given by [22]:
tion from the shearing mechanism to the Orowan strengthening  Gb  d 
mechanism occurs in binary Al–Sc alloys in the particle diameter Or = K4 M(1 − v)−0.5 ln
s
(1)
range from 4 to 6 nm [22,36,37], suggesting that the Orowan  b
Y. Deng et al. / Journal of Alloys and Compounds 530 (2012) 71–80 79

Table 4 Sc was added. In the presence of 0.10 wt.% Zr, antirecrystallized


Microstructural features of Al3 (Sc, Zr) particles in studied aged alloys.
effect of scandium appeared at 0.10 wt.% Sc concentration, and
Alloy Al3 (Sc, Zr) particles enhanced with an increase in the content of scandium. Main aging
precipitates were two kinds of GP zones in under-aged alloys,
dm (nm) N (m−3 ) fv
and ␩ phases in peak-aged alloys, respectively. Small additions
Al–Zn–Mg–0.10Sc–0.10Zr 20 3 × 1020 1 × 10−3
of Sc and Zr did not retard or suppress the formation of aging
Al–Zn–Mg–0.25Sc–0.10Zr 19 5 × 1020 1 × 10−3
precipitates in Al–Zn–Mg–Sc–Zr alloy. In addition, the strengthen-
ing effect from aging precipitates was much stronger than from
dm Sc and Zr microalloying additions in Al–Zn–Mg–Sc–Zr alloys. The
ds = (2)
4 strength increments caused by additions of Sc and Zr were mainly
  0.5  derived from refining strengthening of primary Al3 (Sc, Zr) parti-
1 2 dm
= −1 (3) cles formed during solidification, substructure strengthening and
2 3fV 4 Orowan strengthening caused by the secondary Al3 (Sc, Zr) parti-
where M is the Taylor factor and depends on the texture and the cles formed during homogenization treatment. Moreover, the most
orientation of the tensile axis, v and G are the matrix Poisson’s ratio important strengthening mechanisms of Sc and Zr microalloying
and the shear modulus, respectively, b is the magnitude of the Al in aged Al–Zn–Mg alloys was Orowan strengthening of secondary
matrix Burgers vector, K4 is a constant, the value which depends on Al3 (Sc, Zr) particles.
the particle size and distribution, and ds and  are the mean particle
diameter and an effective inter-particle distance, respectively, both Acknowledgments
on the dislocation slip planes. The values in Table 4 and the follow-
ing values of the parameters of Eq. (1) were used to calculate  Or : This work was financially supported by the National Civilian
K4 = 0.127, M = 3.06, v = 0.331 and G = 27.8 GPa, and b = 0.286 nm, so Matched Project of China (JPPT-115-2-948), the National General
that the Orowan strength increments in Al–Zn–Mg–0.10Sc–0.10Zr Pre-research Project of China (51312010402), Hunan Provincial
and Al–Zn–Mg–0.25Sc–0.10Zr alloys related to the Al3 (Sc, Zr) par- Innovation Foundation for Postgraduate (CX2011B111) and the
ticles were about 50 MPa and 65 MPa, respectively. Graduate Degree Thesis Innovation Foundation of Central South
University (2011ybjz014).
4.4.2. Grain boundary strengthening (Hall–Petch)
The standard Hall–Petch equation, Eq. (4), was employed to
relate the yield strength of the material () to the average grain References
size (d).
[1] B.A. Chen, L. Pan, R.H. Wang, G. Liu, P.M. Cheng, L. Xiao, J. Sun, Materials Science
H–P = 0 + kd−1/2 (4) and Engineering A 530 (2011) 607–617.
[2] Y. Deng, Z.M. Yin, J.Q. Duan, K. Zhao, B. Tang, Z.H. He, Journal of Alloys and
Compounds 517 (2012) 118–126.
In this equation,  0 is the intrinsic resistance of the lattice to
[3] W.H. Yuan, Z.Y. Liang, Materials and Design 32 (2011) 4195–4200.
dislocation motion and k is a parameter that describes the rela- [4] C. Booth-Morrison, D.C. Dunand, Da.N. Seidman, Acta Materialia 59 (2011)
tive strengthening contribution of grain boundaries. A k value of 7029–7042.
0.04 MPa × m1/2 [38,39] for aluminum was used to estimate the [5] W.D. Zhang, Y. Liu, J. Yang, J.Z. Dang, H. Xu, Z.M. Du, Materials Characterization
66 (2012) 104–110.
strengthening due to grain boundaries. [6] A. Bommareddy, M.Z. Quadir, M. Ferry, Journal of Alloys and Compounds (2012),
The increase in the YS due to the grain boundary strengthen- http://dx.doi.org/10.1016/j.jallcom. 2012.02.181.
ing from Sc and Zr microalloying additions,  H–P , was given as [7] A.K. Mukhopadhyay, A. Kumar, S. Raveendra, I. Samajdar, Scripta Materialia 64
(2011) 386–389.
follows: [8] S. Dev, A.A. Stuart, R.C. Ravi Dev Kumaar, B.S. Murty, K.P. Rao, Materials Science
−1/2 −1/2 and Engineering A 467 (2007) 132–138.
H–P = k(dSc-containing − dSc-free ) (5) [9] W. Lefebvre, F. Danoix, H. Hallem, B. Forbord, A. Bostel, K. Marthinsen, Journal
of Alloys and Compounds 470 (2009) 107–110.
By substituting of k value and grain sizes in three stud- [10] N.M. Han, X.M. Zhang, S.D. Liu, D.G. He, R. Zhang, Journal of Alloys and Com-
ied alloys,  H–P can be calculated, which was about pounds 509 (2011) 4138–4145.
[11] M. Vlach, I. Stulíková, B. Smola, N. Zaludová, J. Cerná, Journal of Alloys and
20 MPa in Al–Zn–Mg–0.10Sc–0.10Zr alloy and 28 MPa in Compounds 492 (2010) 143–148.
Al–Zn–Mg–0.25Sc–0.10Zr alloy respectively. [12] D. Geist, S. Ii, K. Tsuchiya, H.P. Karnthaler, G. Stefanov, C. Rentenberger, Journal
Therefore, the calculated YS increases from Sc and Zr of Alloys and Compounds 509 (2011) 1815–1818.
[13] K.E. Knipling, D.N. Seidman, D.C. Dunand, Acta Materialia 59 (2011) 943–
microalloying additions were about 70 MPa and 93 MPa in 954.
Al–Zn–Mg–0.10Sc–0.10Zr alloy and Al–Zn–Mg–0.25Sc–0.10Zr [14] M.E. Van Dalen, T. Gyger, D.C. Dunand, D.N. Seidman, Acta Materialia 59 (2011)
alloy, respectively, which approached the experimental results 7615–7626.
[15] C. Booth-Morrison, D.C. Dunand, DaN. Seidman, Acta Materialia 59 (2011)
(66 MPa and 96 MPa). Moreover, it can be seen that, the most impor-
7029–7042.
tant mechanism from Sc and Zr additions in Al–Zn–Mg–Sc–Zr alloys [16] A. Kverneland, V. Hansen, G. Thorkildsen, H.B. Larsen, P. Pattison, X.Z. Li, J.
was Orowan strengthening of secondary Al3 (Sc, Zr) particles. Gjønnes, Materials Science and Engineering A 528 (2011) 880–887.
[17] X.M. Li, M.J. Starink, Journal of Alloys and Compounds 509 (2011) 471–
476.
5. Conclusions [18] A.R. Eivani, H. Ahmed, J. Zhou, J. Duszczyk, Materials Science and Engineering
A 527 (2010) 2418–2430.
[19] W.T. Wang, X.M. Zhang, Z.G. Gao, Y.Z. Jia, L.Y. Ye, D.W. Zheng, L. Liu, Journal of
Compared with the mechanical properties of peak-aged
Alloys and Compounds 491 (2010) 366–371.
Al–Zn–Mg alloy, the yield strength and the ultimate tensile strength [20] V.G. Davydov, V.I. Elagin, V.V. Zakharov, T.D. Rostova, Metal Science and Heat
increased by 66 MPa and 31 MPa after adding 0.10 wt.% Sc and Treatment 38 (1996) 347–350.
[21] Z.H. Jia, G.Q. Hu, B. Forbord, S.J. Ketil, Materials Science and Engineering A 444
0.10 wt.% Zr, respectively, and improved by 96 MPa and 58 MPa by
(2007) 284–290.
adding 0.25 wt.% Sc and 0.10 wt.% Zr, respectively, meanwhile the [22] O.N. Senkov, M.R. Shagiev, S.V. Senkova, D.B. Miracle, Acta Materialia 56 (2008)
elongations all kept above 12% in two peak-aged Al–Zn–Mg–Sc–Zr 3723–3738.
alloys. In the case of 0.10 wt.% Zr additives in Al–Zn–Mg alloy, mak- [23] L.M. Wu, W.H. Wang, Y.F. Hsub, S. Trong, Journal of Alloys and Compounds 456
(2008) 163–169.
ing 0.25 wt.% Sc scandium additions refined the grains of ingots, [24] K. Yu, W.X. Li, S.R. Li, J. Zhao, Materials Science and Engineering A 368 (2004)
while the modifying effect did not occur when only 0.10 wt.% 88–93.
80 Y. Deng et al. / Journal of Alloys and Compounds 530 (2012) 71–80

[25] K.E. Knipling, R.A. Karnesky, C.P. Lee, D.C. Dunand, D.N. Seidman, Acta Materialia [32] C.B. Fuller, A.R. Krause, D.C. Dunand, D.N. Seidman, Materials Science and Engi-
58 (2010) 5184–5195. neering A 338 (2002) 8–16.
[26] L.M. Wu, M. Seyring, M. Rettenmayr, Materials Science and Engineering A 527 [33] V.G. Davydov, T.D. Rostova, V.V. Zakharov, Yu.A. Filatov, V.I. Yelagin, Materials
(2010) 1068–1073. Science and Engineering A 280 (2000) 30–36.
[27] S. Iwamura, Y. Miura, Acta Materialia 52 (2004) 591–600. [34] O.N. Senkov, R.B. Bhat, S.V. Senkova, J.D. Schloz, Metallurgical and Materials
[28] L.K. Berg, J. Gjønnes, V. Hansen, X.Z. Li, M. Knutson-wedel, G. Waterloo, D. Transactions A 36 (2005) 2115–2126.
Wallenberg, Acta Materialia 49 (2001) 3443–3451. [35] E. Nembach, Particle Strengthening of Metals and Alloys, John Wiley, New York,
[29] X.Z. Li, V. Hansen, J. Gjønnes, L.R. Wallenberg, Acta Materialia 47 (1999) 1997.
2651–2659. [36] K.L. Kendig, D.B. Miracle, Acta Materialia 50 (2002) 4165–4175.
[30] K.B. Hyde, A.F. Norman, P.B. Prangnell, Acta Materialia 49 (2001) [37] C.B. Fuller, D.N. Seidman, D.C. Dunand, Acta Materialia 51 (2003) 4803–4814.
1327–1337. [38] N. Hansen, Scripta Materialia 51 (2004) 801–806.
[31] V.I. Elagin, V.V. Zakharov, T.D. Rostova, Metal Science and Heat Treatment 36 [39] E. Bonetti, L. Pasquini, E. Sampaolesi, Acta Materialia 9 (1997) 611–614.
(1994) 375–380.

You might also like