Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Materials Characterization 102 (2015) 122–130

Contents lists available at ScienceDirect

Materials Characterization
journal homepage: www.elsevier.com/locate/matchar

Effects of two-step homogenization on precipitation behavior of Al3Zr


dispersoids and recrystallization resistance in 7150 aluminum alloy
Zhanying Guo a,b, Gang Zhao b, X.-Grant Chen a,⁎
a
Department of Applied Science, University of Québec at Chicoutimi, Saguenay, QC G7H 2B1, Canada
b
Key Laboratory for Anisotropy and Texture of Materials, Northeastern University, Shenyang 110819, China,

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

Article history: The effect of two-step homogenization treatments on the precipitation behavior of Al3Zr dispersoids was in-
Received 16 December 2014 vestigated by transmission electron microscopy (TEM) in 7150 alloys. Two-step treatments with the first step
Received in revised form 20 February 2015 in the temperature range of 300–400 °C followed by the second step at 470 °C were applied during homogeniza-
Accepted 23 February 2015
tion. Compared with the conventional one-step homogenization, both a finer particle size and a higher number
Available online 25 February 2015
density of Al3Zr dispersoids were obtained with two-step homogenization treatments. The most effective dis-
Keywords:
persoid distribution was attained using the first step held at 300 °C. In addition, the two-step homogenization
7150 aluminum alloy minimized the precipitate free zones and greatly increased the number density of dispersoids near dendrite
Al3Zr dispersoids grain boundaries. The effect of two-step homogenization on recrystallization resistance of 7150 alloys with dif-
Two-step homogenization ferent Zr contents was quantitatively analyzed using the electron backscattered diffraction (EBSD) technique.
Recrystallization It was found that the improved dispersoid distribution through the two-step treatment can effectively inhibit
the recrystallization process during the post-deformation annealing for 7150 alloys containing 0.04–0.09 wt.%
Zr, resulting in a remarkable reduction of the volume fraction and grain size of recrystallization grains.
© 2015 Elsevier Inc. All rights reserved.

1. Introduction The effects of both the Zr content and the homogenization condition
on the precipitation of Al3Zr dispersoids and the recrystallization of Al
7xxx series aluminum alloys have high strength to density ratio and alloys have been investigated in previous studies [14,16–20]. The con-
excellent mechanical properties, such as high strength, high fracture ventional homogenization treatment (one-step) is mainly intended to
toughness and good resistance to stress corrosion, and are widely dissolve low melting eutectic phases and to redistribute the solutes in
used in the aeronautical and astronautic industries [1–4]. However, re- aluminum matrix, but not necessary to optimize the precipitation of
crystallization can significantly deteriorate mechanical properties of Al3Zr dispersoids. The high homogenization temperature of the conven-
heat-treated materials; moreover, the fracture toughness of 7xxx alloys tional practice causes both high solubility and diffusion rate of zirconi-
decreases with the increase of the recrystallization degree that occurs um, which results in a low volume fraction of large dispersoids and
during the hot-forming process and solution heat treatment [4]. wide precipitate free zones. To improve the distribution of Al3Zr disper-
Small quantities of Zr are commonly added into 7xxx Al alloys as a soids, some approaches to stepwise homogenization were explored in
recrystallization inhibitor, in which thermally stable Al3Zr dispersoids the literature [12,19–22]. The first step was used to create a favorable
precipitate during the homogenization of cast ingots. Al alloys with Zr condition of dispersoid precipitation for developing an optimized distri-
additions can retain their deformed structure and prevent recrystalliza- bution. The subsequent step then completed the homogenization of the
tion, due to the presence of Al3Zr dispersoids, even when annealed at alloy in the conventional manner. However, the nucleation, growth and
high temperature [5]. The effectiveness of Al3Zr dispersoids strongly de- distribution of dispersoids during the homogenization process are not
pends on their size, number density and distribution [6]. It is recognized well understood and the related recrystallization behavior of deformed
that Zr tends to segregate and enrich in the dendrite centers during materials was not quantitatively examined.
solidification [7–9]. It is difficult to remove Zr concentration gradients In the present study, the effects of one-step and various two-step ho-
during conventional homogenization due to Zr's low diffusivity in Al mogenization treatments on the precipitation behavior of Al3Zr disper-
[10,11]. Therefore, the precipitation of Al3Zr dispersoids often concen- soids in the 7150 alloy were investigated. Particular attention was paid
trates in the center of dendrite grains and the precipitate free zone to the formation of Al3Zr dispersoids during the first step of homogeni-
(PFZ) of dispersoids forms in dendrite grain boundaries [12–15]. zation, which was designed to promote the nucleation of dispersoids
and thus to improve the overall distribution of Al3Zr dispersoids in the
⁎ Corresponding author. final homogenization step. The influence of two-step homogeniza-
E-mail address: [email protected] (X.-G. Chen). tion conditions on the recrystallization resistance of 7150 alloys

http://dx.doi.org/10.1016/j.matchar.2015.02.016
1044-5803/© 2015 Elsevier Inc. All rights reserved.
Z. Guo et al. / Materials Characterization 102 (2015) 122–130 123

with varying Zr content was quantitatively analyzed during the Table 2


post-deformation annealing. Temperature and time of one-step and two-step homogenization treatments applied.

Heat treatment First step Second step

Temperature (°C) Time (h) Temperature (°C) Time (h)


2. Experimental procedure
One-step – – 470 24
Two-step 300 48 470 24
Four experimental alloys with different Zr contents were prepared 300 72 470 24
by the ingot metallurgical route. The alloys were melted and batched 400 48 470 24
in a graphite crucible using an electrical resistance furnace. The melt 400 72 470 24
was poured at a temperature of 750 °C into a rectangular permanent
steel mold to produce cast ingots measuring 30 × 40 × 80 mm3. The
chemical compositions of the experimental alloys are listed in Table 1 the center of each sample were selected for the recrystallization
(all compositions are in wt.% unless otherwise indicated). The homoge- analysis.
nization processes included both conventional homogenization (one-
step) where samples were held at 470 °C for 24 h and the new two- 3. Results and discussion
step homogenization treatments. For the two-step homogenization,
two temperatures (300 °C and 400 °C) and two holding times (48 h 3.1. Precipitation behavior of Al3Zr dispersoids during homogenization
and 72 h) were used in the first step of treatment, while the second
step was the same as the one-step homogenization. After the comple- The study of the precipitation of the Al3Zr dispersoids during
tion of the homogenization treatment, the samples were subjected to homogenization was focused on Alloy C (0.09% Zr). Typical TEM dark
direct water quenching. To study the effect of each step of homogeniza- field images of the dispersoid precipitation after both one-step and
tion on the dispersoid precipitation, some selected samples after the two-step homogenization treatments are presented in Fig. 1. The corre-
first step of treatment were also water-quenched. The details of both sponding selected area diffraction (SAD) pattern, as shown in the inset
the one-step and two-step homogenization treatments are provided of Fig. 1a, indicated that the precipitates were Al3Zr with a L12 crystal
in Table 2. structure. Compared with the one-step homogenization, both a finer
Cylindrical compression samples 10 mm in diameter and 15 mm in particle size and a higher number density of Al3Zr dispersoids are
height were machined from the homogenized samples. Uniaxial hot com- obtained for all two-step homogenization conditions. The average radi-
pression tests were performed on a Gleeble 3800 thermomechanical sim- us, number density and volume fraction of dispersoids are listed in
ulator at a deformation temperature of 400 °C with a constant strain rate Table 3. It can be seen from Fig. 1 and Table 3 that the two-step homog-
of 1 s−1. Compression samples were deformed to a total true strain of 0.8 enization with first step held at 300 °C has the most significant impact
and subsequently water-quenched. To evaluate the recrystallization resis- on the precipitation of Al3Zr dispersoids, resulting in the finest particle
tance after the post-deformation heat treatment, the deformed samples size and densest distribution of dispersoids (Fig. 1b and c). Results of
were annealed at 470 °C for 2 h, followed by water quenching to room two holding times (48 and 72 h) are quite similar, and the dispersoid
temperature. size and number density change only slightly when the holding time
To study the precipitation of Al3Zr dispersoids, the homogenized is prolonged from 48 to 72 h at 300 °C. On the other hand, the two-
samples were examined under a transmission electron microscope step homogenization with the first step treated at 400 °C (Fig. 1d and
(TEM, JEOL JEM-2100) operated at 200 kV. TEM specimens were f) also results in a considerable reduction of dispersoid size (14 nm vs.
mechanically ground to approximately 40-μm thickness, followed by 20.6 nm of the one-step homogenization). However, prolonging the
twin-jet electropolishing at 15 V DC in a 30% nitric acid and 70% meth- holding time from 48 to 72 h at 400 °C causes a slight increase in particle
anol solution cooled to −25 °C. The TEM examination was performed size and a minor decrease in the number density of dispersoids. With re-
with the specimen oriented along low index [011] zone axis of Al spect to the volume fraction of dispersoids, the values of the volume
matrix, utilizing two-beam diffraction conditions. Centered dark field fraction are nearly constant for all homogenization conditions, indicat-
! ing that all supersaturated Zr solutes are out of the solution to form
images of the Al3Zr dispersoids were formed using a g = (200) L12
superlattice reflection by tilting the incident illumination by an angle dispersoids and the volume fraction of Al3Zr dispersoids reaches its
equal to the diffraction angle. To determine the particle volume fraction, equilibrium value after both one-step and two-step treatments.
the thickness of the TEM foils was measured using an Electro Energy Fig. 2 shows TEM dark field images of Al3Zr precipitation with the
Loss Spectroscopy (EELS) attached in the TEM. The average radius, first step held at 300 °C. After being held for 48 h, some fine dispersoids
number density and volume fraction of the dispersoids were quantified can be observed but the number of the dispersoids seems to be much
by image analysis of digitized TEM dark field images. 10 different areas smaller than that after the two-step homogenization (Fig. 2a vs. Fig. 1b).
(1.15 × 1.15 μm2 for each) in the dendrite center and more than 200 By careful observation at high magnification (Fig. 2b), it is found that,
particles were measured for each homogenization condition. apart from relatively large dispersoids, many small dispersoids with a di-
To quantify the recrystallized structures, all annealed samples were ameter of approximately 1–3 nm also appeared in the aluminum matrix.
sectioned and polished parallel to the compression axis along the cen- Those small dispersoids are marked with arrows and their quantities are
terline and then examined using the electron backscattered diffraction few times larger than the relatively large dispersoids. It should be men-
(EBSD) technique under a SEM (JEOL JSM-6480LV). The surface scan- tioned that the conventional TEM used in this study has a difficulty in
ning area of 0.8 mm2 with a scanning step size of 3 μm and 3 areas at clearly revealing the dispersoids smaller than 1–1.5 nm. It is reasonable
to believe that there are a certain number of smaller dispersoids that
could not fully be revealed. It is evident that the first step treatment
at 300 °C promotes the nucleation of a large number of dispersoids
Table 1
Chemical compositions of experimental alloys (wt.%).
out of the aluminum matrix due to a high Zr supersaturation at
low temperature. Many of those small dispersoids can only slowly
Alloys Zn Mg Cu Zr Fe Si Ti Al grow because of the low diffusion rate of Zr atoms at the low tem-
A (base) 6.15 2.10 2.11 0 0.13 0.11 0.008 Bal. perature of 300 °C (6.34 × 10− 24 m2 s− 1). The diffusion rate of Zr
B 6.10 2.00 2.15 0.04 0.13 0.11 0.008 Bal.
C 5.99 1.91 2.10 0.09 0.13 0.11 0.008 Bal.
in Al matrix at a given temperature is calculated from D ¼ D0 exp
 
D 6.00 2.08 2.09 0.16 0.13 0.12 0.008 Bal. ‐Q −2
RT , where D0 = 7.28 × 10 m2 s− 1 and Q = 242 kJ mol -1 [10].
124 Z. Guo et al. / Materials Characterization 102 (2015) 122–130

(a)

(b) (c)

(d) (e)

Fig. 1. Typical TEM centered dark field images of Al3Zr dispersoids in 7150–0.09Zr alloy after both one-step and two-step homogenization treatments: (a) 470 °C/24 h with an inset of SAD
pattern, (b) 300 °C/48 h + 470 °C/24 h, (c) 300 °C/72 h + 470 °C/24 h, (d) 400 °C/48 h + 470 °C/24 h and (e) 400 °C/72 h + 470 °C/24 h.

The diffusion rates of Zr in Al matrix at the studied temperatures of 300, precipitates are also verified being Al3Zr by the SAD pattern, as pre-
400 and 470 °C are 6.34 × 10− 24 , 1.20 × 10− 20 and 7.07 × sented in the inset of Fig. 3a. However, all Al3Zr dispersoids already
10− 19 m2 s− 1, respectively. grow to a relatively large size (Fig. 3a) due to a high diffusion rate of
After the first step treated at 400 °C, the number of dispersoids is Zr at 400 °C (1.20 × 10− 20 m2 s− 1), and it is difficult to find small
clearly increased relative to the one-step homogenization. The dispersoids between the existing ones (Fig. 3b). The number of
Z. Guo et al. / Materials Characterization 102 (2015) 122–130 125

Table 3
Average radius, number density and volume fraction of Al3Zr dispersoids in 7150–0.09Zr alloy after homogenization treatments.

Homogenization conditions Homogenization parameters Average radius Number density Volume fraction f/r
r (nm) N (μm−3) f (%) (μm−1)

One-step 470 °C/24 h 20.6 85 0.342 0.164


Two-step 300 °C/48 h + 470 °C/24 h 10.7 629 0.338 0.314
300 °C/72 h + 470 °C/24 h 10.3 673 0.333 0.323
400 °C/48 h + 470 °C/24 h 13.4 290 0.332 0.248
400 °C/72 h + 470 °C/24 h 13.9 279 0.339 0.243

dispersoids observed after first step treated at 400 °C is almost iden- the first step treated at lower temperatures is considerably higher be-
tical to that after the completed two-step treatment. cause the Zr supersaturation is largely increased at lower temperature.
During homogenization, there are two main factors influencing the Therefore, a larger number of dispersoids can be formed by the first
dispersoid precipitation: (1) the supersaturation of the solid solution step treatment. Moreover, the lower the temperature at the first step,
(driving force for nucleation) and (2) the diffusivity of the solute in the higher the driving force for the nucleation but the lower the growth
the matrix (for growth and coarsening). When treated at the high tem- rate of the dispersoids are. The first step treatment at 300 °C produced
perature of the conventional one-step practice, fewer dispersoids can more Al3Zr nuclei than that at 400 °C. However, those small dispersoids
form due to a low supersaturation of Zr in solid solution but they can only slowly grow at this temperature. The subsequent second step
grow faster and may also coarsen because of the high diffusion rate of treatment provides a normal condition for their growth. Overall, as
Zr solutes at 470 °C (7.07 × 10−19 m2 s−1). Compared with the one- shown in Table 3, much denser and finer dispersoids were obtained
step homogenization, the driving force for dispersoid nucleation by when the first step temperature was reduced from 400 to 300 °C.

(a) (b)

Area A

Fig. 2. TEM centered dark field images of Al3Zr dispersoids precipitated after the first step treated at (a) 300 °C for 48 h and (b) enlarged area A.

(a) (b)
Area B

Fig. 3. TEM centered dark field images of Al3Zr dispersoids precipitated after the first step treated at (a) 400 °C for 48 h with an inset of SAD pattern and (b) enlarged area B.
126 Z. Guo et al. / Materials Characterization 102 (2015) 122–130

(a)

Grain boundary

1u m

(b)

Grain boundary

1u m

Fig. 4. TEM dark field images of the Al3Zr dispersoid distribution in grain boundary regions, subjected to different homogenization treatments: (a) one-step, 470 °C/24 h, (b) two-step,
300 °C/48 h + 470 °C/24 h.

The precipitate free zone (PFZ) of Al3Zr dispersoids in the dendrite interdendritic region. Therefore, the precipitation of Al3Zr dispersoids
grain boundary can be clearly observed after one-step homogenization principally occurs in the dendrite cores and the precipitate free zone
at 470 °C (Fig. 4a). Having a solid–liquid partition coefficient, k0 = 2.5, (PFZ) of Al3Zr dispersoids forms near the interdendritic grain boundary
larger than unity, Zr as a peritectic element segregates inversely from during homogenization. The widths of PFZ are in the range of 2 to 3 μm
the dendrite center to the interdentritic boundary during solidification, and the distribution of dispersoids near PFZ is non-uniform. Close to the
resulting in solute-rich dendritic cores surrounded by solute-depleted boundary, only few dispersoids with large interparticle space can be

(a) (b)

200um 200um

(c) (d

200um 200um

Fig. 5. Orientation maps of 7150 alloys with different Zr contents for one-step homogenization: (a) the base alloy (0% Zr); (b) 0.04% Zr; (c) 0.09% Zr; (d) 0.16% Zr. High angle boundaries
(N15°) and low angle boundaries (1–15°) shown as black line and white line, respectively.
Z. Guo et al. / Materials Characterization 102 (2015) 122–130 127

the dendrite grain boundary, leading to a narrow PFZ and a dense distri-
bution of dispersoids after two-step homogenization.

3.2. Effect of Zr contents and homogenization treatments on recrystallization


resistance

In the 7xxx wrought alloys, recrystallization during post-deformation


heat treatments could considerably deteriorate the alloy strength and
fracture toughness. To study the effect of Zr on the recrystallization
behavior of the 7150 alloys during post-deformation heat treatments,
hot-deformed samples were isothermally annealed at 470 °C for 2 h
and subsequently water-quenched. The influence of the Zr contents on
the recrystallization with the one-step homogenization is shown in
Fig. 5. Samples with Zr contents from 0 to 0.09% show all partially re-
crystallized microstructures after annealing, and the sample with 0.16%
Zr exhibits a main recovery microstructure with few, small and isolated
Fig. 6. Volume fraction and average size of recrystallized grains as a function of Zr contents recrystallized grains. During annealing, recrystallized grains with high-
for one-step homogenization. angle boundaries (N15°) were developed in those alloys, indicating the
occurrence of static recrystallization. The volume fraction and average
seen. Towards the dendrite center, the number of dispersoids is in- size of the recrystallized grains were quantitatively analyzed using the
creased and their distribution becomes more homogeneous, which is EBSD technique, and the results are plotted in Fig. 6. The determination
consistent with previous studies [14,22]. Compared with the one-step of statically recrystallized grains is based on two criteria [23]: (1) the
homogenization, the widths of PFZ after two-step homogenization misorientation of the boundary between the recrystallized grain and the
with first step treated at 300 °C are reduced from 2–3 μm to 1–1.5 μm. deformed matrix is larger than 15°; and (2) the mean misorientation
Moreover, the number of dispersoids close to the boundary increases within the recrystallized grain is lower than 1°. As shown in Fig. 6, the vol-
significantly and the size of the dispersoids decreases remarkably ume fraction and average grain size of recrystallized grain decrease with
(Fig. 4b). increasing Zr content. The recrystallized fraction decreases from 75% of
During non-equilibrium solidification, the local Zr concentration Alloy A (base alloy without Zr) down to 1% of Alloy D (0.16% Zr), indicat-
gradually decreases from the dendrite center to the boundary [8]. The ing a significant increase in recrystallization resistance when Zr content
one-step homogenization at a relatively high temperature (470 °C) re- increases.
sults in an insufficient Zr supersaturation close to the boundary for the The microstructure evolution of Alloy B (0.04% Zr) subject to two-
dispersoid precipitation. With the first step treated at lower tempera- step homogenization treatments is shown in Fig. 7. All samples are par-
ture, the Zr supersaturation in the same region is large and hence the tially recrystallized after annealing. The measured recrystallized fraction
driving force for the nucleation of the dispersoids is significantly en- and grain size for both one-step and two-step homogenization treat-
hanced. Therefore, a larger number of dispersoids can be formed near ments are shown in Fig. 8. Compared with the sample with the one-step

(a) (b)

200um 200um

(c) (d)

200um 200um

Fig. 7. Orientation maps of 7150–0.04% Zr alloy homogenized at: (a) 300 °C/48 h + 470 °C/24 h; (b) 300 °C/72 h + 470 °C/24 h; (c) 400 °C/48 h + 470 °C/24 h; (d) 400 °C/72 h + 470 °C/24 h.
128 Z. Guo et al. / Materials Characterization 102 (2015) 122–130

Coherent Al3Zr dispersoids are known to effectively prevent the mo-


tion of subgrain boundaries during annealing, hence retarding the static
recrystallization process. The effectiveness of dispersoids in preventing
recrystallization can be quantified by calculating the Zener pinning
pressure, PZ, in the following equation [6,24,25]:

3γGB
PZ ¼ ð f =r Þ ð1Þ
2

where γGB is the specific grain boundary energy, f is the volume fraction
and r is the average radius of dispersoids. Eq. (1) indicates that a high
volume fraction of small dispersoids (i.e., a high f/r value) is necessary
to achieve a high PZ on grain boundary migration for retarding the
growth of recrystallized grains. PZ can be greatly increased by maximiz-
ing the volume fraction (f) and minimizing the dispersoid size (r). Com-
pared with the one-step homogenization, higher f/r values of Al3Zr
dispersoids are obtained when two-step homogenization treatments
are applied (Table 3), particularly for the first step treated at low tem-
perature (300 °C). The f/r value increases from 0.16 of the one-step ho-
mogenization sample to 0.32 of the two-step homogenization samples
with first step treated at 300 °C. In the alloys subject to two-step homog-
enization treatments, a great number of fine Al3Zr dispersoids dispersed
in the aluminum matrix, and they played a major role in inhibiting re-
crystallization. On the other hand, the recrystallization often occurs
near dendrite grain boundaries where few dispersoids are present and
the PFZ locates (a low local f/r value). The two-step homogenization
treatment can reduce the width of PFZ and increase the number of dis-
persoids near dendrite grain boundaries (Fig. 4), leading to a high local
f/r value and hence an increased PZ in the region. As a consequence,
samples subject to two-step homogenization treatments display a re-
markably high recrystallization resistance. The highest recrystallization
resistance is obtained when first step treated at 300 °C due to the
highest number density and finest size of dispersoids.

4. Conclusions

Fig. 8. Volume fraction and average size of recrystallized grain of 7150–0.04% Zr alloy for (1) The conventional one-step homogenization used for the 7150
both one-step and two-step homogenization treatments: (a) volume fraction and alloy produced a low number density and non-uniform dis-
(b) average grain size.
tribution of Al3Zr dispersoids. Compared with the one-step ho-
mogenization treatment, both a finer particle size and a higher
number density of Al3Zr dispersoids were obtained with the
homogenization, the recrystallized fractions of the two-step homogeniza- two-step homogenization treatments. When the first step was
tion samples are generally reduced, indicating a clear increase in the treating at 300 and 400 °C, the number density of dispersoids in-
recrystallization resistance of the samples with two-step homogenization creased by 7–8 and 3–3.5 times, respectively. The most effective
treatments. Moreover, the samples with the first step treated at 300 °C dispersoid distribution was attained using the first step held at
show more obvious effects that that treated at 400 °C. Both the recrystal- 300 °C.
lized fraction and grain size with first step treated at 300 °C are remark- (2) The two-step homogenization treatment minimized the precipi-
ably reduced relative to the one-step homogenization sample. Results tate free zones and greatly increased the number density of dis-
also show that the holding time for the 300 °C and 400 °C conditions persoids near dendrite grain boundaries, leading to a significant
has a negligible impact on recrystallized fraction and grain size of recrys- improvement of dispersoid distribution in aluminum matrix.
tallization grains. (3) The improved dispersoid distribution through the two-step ho-
The microstructure evolution of Alloy C (0.09% Zr) with the two-step mogenization treatments can effectively inhibit the recrystalliza-
homogenization treatments is shown in Fig. 9. When the samples are tion process during post-deformation annealing for 7150 alloys
subject to two-step homogenization treatments, their microstructure containing 0.04–0.09% Zr. Compared with the one-step homoge-
becomes a main recovery structure with few isolated recrystallized nization, both volume fraction and grain size of recrystallization
grains emerging. Compared with the sample subject to the one-step grains with two-step homogenization treatments were remark-
homogenization that is partially recrystallized (Fig. 5c), the recrystalli- ably reduced. The first step treated at low temperature (300 °C)
zation process is almost inhibited with two-step homogenization was found to be the most effective to increase the recrystalliza-
treatments. The measured recrystallized fraction and grain size are tion resistance.
also displayed in Fig. 10. The recrystallized fraction and grain size are re-
duced to approximately 2% and 25 μm, respectively, for both first steps
treated at 300 °C and 400 °C. For the sample of Alloy C (0.16% Zr), the Acknowledgments
recrystallization resistance of the alloy is already very high with the
one-step homogenization and the recrystallized fraction is only approx- The authors would like to acknowledge the financial support from the
imately 1% (Figs. 5d and 6). Therefore, no two-step homogenization Natural Sciences and Engineering Research Council of Canada (NSERC)
treatments were applied. and from Rio Tinto Alcan through the NSERC Industrial Research Chair
Z. Guo et al. / Materials Characterization 102 (2015) 122–130 129

(a) (b)

200um 200um

(c) (d)

200um 200um

Fig. 9. Orientation maps of 7150–0.09% Zr alloy homogenized at: (a) 300 °C/48 h + 470 °C/24 h; (b) 300 °C/72 h + 470 °C/24 h; (c) 400 °C/48 h + 470 °C/24 h; (d) 400 °C/72 h + 470 °C/24 h.

in Metallurgy of Aluminum Transformation at the University of Québec at [13] Z.H. Jia, G.H. Hu, B. Forbord, J.K. Solberg, Effect of homogenization and alloying
elements on recrystallization resistance of Al–Zr–Mn alloys, Mater. Sci. Eng. A 444
Chicoutimi. The authors would also like to thank Dr. Z. Zhang for his help (2007) 284–290.
in TEM observation. [14] J.D. Robson, P.B. Prangnell, Predicting recrystallised volume fraction in aluminium
alloy 7050 hot rolled plate, Mater. Sci. Technol. 18 (2002) 607–614.
[15] Y.L. Deng, L. Wan, L.H. Wu, Y.Y. Zhang, X.M. Zhang, Microstructural evolution of
References Al–Zn–Mg–Cu alloy during homogenization, J. Mater. Sci. 46 (2011) 875–881.
[16] B. Morere, R. Shahani, C. Maurice, J. Driver, The influence of Al3Zr dispersoids on the
[1] J.P. Immarigeon, R.T. Holt, A.K. Koul, L. Zhao, W. Wallace, J.C. Beddoes, Light weight
recrystallization of hot-deformed AA 7010 alloys, Metall. Mater. Trans. A 32 (2001)
materials for aircraft applications, Mater. Charact. 35 (1995) 41–67.
625–632.
[2] A. Heinz, A. Haszler, C. Keidel, S. Moldenhauer, R. Benedictus, W.S. Miller, Recent
[17] J.D. Robson, P.B. Prangnell, Modelling Al3Zr dispersoid precipitation in multicomponent
development in aluminium alloys for aerospace applications, Mater. Sci. Eng. A
aluminium alloys, Mater. Sci. Eng. A 352 (2003) 240–250.
280 (2000) 102–107.
[18] A.R. Eivani, H. Ahmed, J. Zhou, J. Duszczyk, An experimental and theoretical investiga-
[3] J.C. Williams, E.A. Starke, Progress in structural materials for aerospace systems, Acta
tion of the formation of Zr-containing dispersoids in Al–4.5Zn–1 Mg aluminum alloy,
Mater. 51 (2003) 5775–5799.
Mater. Sci. Eng. A 527 (2010) 2418–2430.
[4] E.A. Starke, J.T. Staley, Application of modern aluminum alloys to aircraft, Prog.
[19] Z.H. Jia, G.H. Hu, B. Forbord, J.K. Solberg, Enhancement of recrystallization resistance
Aerosp. Sci. 32 (1996) 131–172.
of Al–Zr–Mn by two-step precipitation annealing, Mater. Sci. Eng. A 483–484 (2008)
[5] V. Ocenasek, M. Slamova, Resistance to recrystallization due to Sc and Zr addition to
195–198.
Al–Mg alloys, Mater. Charact. 47 (2001) 157–162.
[20] X.Y. Lu, E.J. Guo, P. Rometsch, L.J. Wang, Effect of one-step and two-step homogeniza-
[6] F.J. Humphreys, M. Hatherly, Recrystallization and Related Annealing Phenomena,
tion treatments on distribution of Al3Zr dispersoids in commercial AA7150 aluminium
3rd ed. Elsevier Science Inc., Oxford, 1995.
alloy, Trans. Nonferrous Met. Soc. 22 (2012) 2645–2651.
[7] K.E. Knipling, D.C. Dunand, D.N. Seidman, Nucleation and precipitation strengthening in
[21] B.L. Ou, J.G. Yang, M.Y. Wei, Effect of homogenization and aging treatment on
dilute Al–Ti and Al–Zr alloys, Metall. Mater. Trans. A 38 (2007) 2552–2563.
mechanical properties and stress-corrosion cracking of 7050 alloys, Metall. Mater.
[8] K.E. Knipling, D.C. Dunand, D.N. Seidman, Precipitation evolution in Al–Zr and Al–Zr–Ti
Trans. A 38 (2007) 1760–1773.
alloys during isothermal aging at 375–425 oC, Acta Mater. 56 (2008) 114–127.
[22] Y.L. Deng, Y.Y. Zhang, L. Wan, A.A. Zhu, X.M. Zhang, Three-stage homogenization
[9] K.E. Knipling, D.C. Dunand, D.N. Seidman, Ambient- and high-temperature mechanical
of Al–Zn–Mg–Cu alloys containing trace Zr, Metall. Mater. Trans. A 44 (2013)
properties of isochronally aged Al–0.06Sc, Al–0.06Zr and Al–0.06Sc–0.06Zr (at.%) alloys,
2470–2477.
Acta Mater. 59 (2011) 943–954.
[23] F.J. Humphreys, Review—grain and subgrain characterisation by electron backscatter
[10] E. Clouet, J.M. Sanchez, C. Sigli, First-principles study of the solubility of Zr in Al,
diffraction, J. Mater. Sci. 36 (2001) 3833–3854.
Phys. Rev. B 65 (2002) 094105.
[24] E. Nes, N. Ryum, O. Hunderi, On the Zener drag, Acta Metall. Mater. 33 (1985) 11–22.
[11] M. Schobel, P. Pongratz, H.P. Degischer, Coherency loss of Al3(Sc, Zr) precipitates by
[25] R.D. Doherty, Role of interfaces in kinetics of internal shape changes, Mater. Sci. 16
deformation of an Al–Zn–Mg alloy, Acta Mater. 60 (2012) 4247–4254.
(1982) 1–13.
[12] J.D. Robson, Optimizing the homogenization of zirconium containing commercial alu-
minium alloys using a novel process model, Mater. Sci. Eng. A 338 (2002) 219–229.
130 Z. Guo et al. / Materials Characterization 102 (2015) 122–130

Fig. 10. Volume fraction and average size of recrystallized grains of 7150–0.09% Zr Al alloy
for both one-step and two-step homogenization treatments: (a) volume fraction and
(b) average grain size.

You might also like