You are on page 1of 7

Materials Science & Engineering A 639 (2015) 307–313

Contents lists available at ScienceDirect

Materials Science & Engineering A


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

A study of precipitation strengthening and recrystallization behavior


in dilute Al–Er–Hf–Zr alloys
H. Wu, S.P. Wen n, X.L. Wu, K.Y. Gao, H. Huang, W. Wang, Z.R. Nie n
School of Materials Science and Engineering, Beijing University of Technology, Beijing 100124, People's Republic of China

art ic l e i nf o a b s t r a c t

Article history:
Received 20 March 2015 A study on the precipitation hardening and recrystallization behavior of dilute Al–Er–Hf and Al–Er–Hf–Zr
Received in revised form
alloys has been carried out. The results show that both Al–0.045Er–0.18Hf and Al–0.045Er–0.08Zr–0.1Hf
6 May 2015
alloys can obtain remarkable age strengthening effect and recrystallization resistance. The precipitation
Accepted 8 May 2015
Available online 16 May 2015 hardening rate of Al–0.045Er–0.08Zr–0.1Hf is accelerated compared with that of the Al–0.045Er–0.18Hf
alloy due to substituting 0.08 at% Zr for Hf, which can be ascribed to the sequential precipitation of solute
Keywords: elements on the basis of the disparity in their intrinsic diffusivities (DEr 4DZr 4DHf). The peak hardness
Aluminum alloy
values for the Al–0.045Er–0.08Zr–0.1Hf are 644 MPa and 662 MPa after isochronal aging to 450 °C and
Precipitation
isothermal aging at 350 °C for 84 h, respectively, which are higher than those of the Al–0.045Er–0.18Hf
Erbium
Hafnium alloy. The recrystallization temperature of Al–Er–Hf–Zr alloy is 450 °C, about 25 °C higher than that of the
Zirconium Al–Er–Hf alloy due to the larger f/r ratio of precipitates in Al–Er–Hf–Zr alloys.
Recrystallization & 2015 Elsevier B.V. All rights reserved.

1. Introduction and high thermal stability are relatively easy to obtain in aluminum
alloys with joint addition of the above mentioned elements. For
Dispersoid phases, especially the thermally stable L12-structured example, Al–Er–Hf and Al–Er–Zr ternary alloys have a pronounced
precipitates are effective to stabilize the sub-grain structure, inhibit strengthening effect and recrystallization resistance due to the
recrystallization and the corresponding strength loss by exerting a higher number density of thermal stable precipitation compared
retarding force on migrating dislocation and grain boundaries with that of the binary Al–Zr, Al–Er and Al–Hf alloys [4,8,11,16]. The
during thermo-mechanical processing and subsequent annealing synergy effects of joint addition of elements in Al–Sc–Hf [17,18] and
[1]. However, only a limited number of alloying elements, such as Al–Sc–Zr [5,19–21] alloys are similar to that of Al–Er–Zr/Hf alloys.
Er, Sc, Yb, Zr and Hf, have the capability to crystallize to form the Furthermore, Hallam's investigations have shown that an even
coherent strengthening phases (Al3M) with ordered L12-structured higher precipitation strengthening and recrystallization resistance
in Al-based alloys [2]. Furthermore, both fast precipitation and high can be obtained in Al–Sc–Zr–Hf quaternary alloy than that in
thermal stability of the precipitates cannot be obtained by a single ternary Al–Sc–Hf alloys [18,22]. In our previous investigation, the
addition of these elements, because both precipitation and coar- Al–Er–Hf alloy has shown an outstanding precipitation hardening
sening are closely related to the diffusivity of these elements in [8]. It is expected to further enhance the strength of this alloy by
aluminum. The nucleation kinetics of L12-structured Al3Hf in Al–Hf addition of quaternary elements. Moreover, substitution of more
alloys is very slow at relatively low temperature, but the low dif- expensive Hf by other elements, such as Zr, will reduce the alloy
fusivity of Hf ensures that Al3Hf particles coarsen slowly at elevated cost. Therefore, in this article, isochronal and isothermal annealing
temperature, and so does in the binary Al–Zr alloys [3–8]. The treatments for Al–Er–Hf–(Zr) alloys have been carried out in order
addition of other elements, such as Sc [5,9,10], Er [4,9,11,12] and Yb to describe the precipitation strengthening effect of Al–Er–Hf–Zr
[9,13–15] leads to rapid nucleation of coherent L12-structured Al3M alloys. In addition, the subsequent effect of the precipitates on
(M¼Er,Yb and Sc) precipitates, but the relatively higher diffusivity recrystallization behavior of the cold rolled alloy has also been
of these element limits the thermal stability of the precipitates at examined.
elevated temperature.
Fortunately, it has been reported that the coherent ordered
L12-structured precipitates with both rapid precipitation kinetics 2. Experimental methods

n
Corresponding authors. Tel./fax: þ86 10 67391536. A ternary and a quaternary alloys with nominal compositions of
E-mail addresses: wensp@bjut.edu.cn (S.P. Wen), zrnie@bjut.edu.cn (Z.R. Nie). Al–0.045Er–0.18Hf and Al–0.045Er–0.1Hf–0.08Zr (with the same

http://dx.doi.org/10.1016/j.msea.2015.05.027
0921-5093/& 2015 Elsevier B.V. All rights reserved.
308 H. Wu et al. / Materials Science & Engineering A 639 (2015) 307–313

total solute concentration), respectively, were prepared by ingot precipitation hardening rate of Al–0.045Er–0.08Zr–0.1Hf is much
metallurgy. Their compositions in the as-cast state are given in faster than that of Al–0.045–0.18Hf at temperature ranging from
Table 1 (all compositions are in at% unless otherwise noted), as 300 °C to 400 °C. The hardness of Al–0.045Er–0.08Zr–0.1Hf
verified by inductively coupled plasma-atomic emission spectro- approaches 525 MPa at 400 °C, which is significantly larger than
scopy. The alloy were dilution casted from the commercially pure that of the Al–0.045–0.18Hf alloy at the same temperature. Then
aluminum (Fe: 0.04 wt%; Si: 0.03 wt%), with Al–6 wt% Er, Al–4 wt% Al–0.045Er–0.08Zr–0.1Hf alloy keep the same precipitation hard-
Hf and Al–4 wt% Zr master alloys. Appropriate amounts of the ening rate as that of the Al–0.045Er–0.18Hf alloy, and obtain a
starting materials were melted in a crucible furnace in a resistively peak hardness value of 644 MPa at 450 °C. It can be found that the
heated furnace to 790 °C. After thorough stirring, the melt was value of peak hardness is about 140 MPa and 40 MPa higher than
poured into an iron mold and quickly cooled to ambient tem-
that of the Al–0.04Er–0.08Zr and Al–0.045Er–0.18Hf alloys,
perature. In order to minimize or eliminate dendrite segregation
respectively. The temperature to obtain the peak hardness is about
during solidification, the castings were homogenized in air at
25 °C lower and higher than that of the Al–0.045Er–0.18Hf and the
640 °C for 24 h and followed by water quenching. The as-solution
Al–0.04Er–0.08Zr alloys, respectively.
alloys were immediately aged isochronally from 150 to 600 °C
Vickers microhardness evolutions of the experimental alloys
with increments of 25 °C, lasting for 3 h at each temperature.
during isothermal aging at 350 °C are displayed in Fig. 2. Data for
Some samples were isothermally aged at 350 °C for different times
ranging from 0.1 to 500 h. In order to evaluate the effect of pre- Al–0.04Er–0.08Zr [4] alloy are presented for comparison. The peak
cipitates on recrystallization resistance of alloys, the isothermally hardness of the Al–0.04Er–0.08Zr, Al–0.045Er–0.18Hf and Al–
peak aged samples were cold rolled 60% to a thickness of 0.045Er–0.08Zr–0.1Hf is 560 MPa, 640 MPa and 660 MPa, respec-
approximately 6 mm. The cold rolled samples were cut parallel to tively. The time to obtain the peak hardness of the Al–0.04Er–
the RD/ND plane (longitudinal section) of the rolled plate and 0.08Zr, Al–0.045Er–0.18Hf and Al–0.045Er–0.08Zr–0.1Hf is 64 h,
annealed in the temperature range of 175–600 °C with increment 84 h and 100 h, respectively. From these results, it can be found
of 25 °C, and at each temperature point the samples were iso- that addition of 0.1Hf in Al–0.04Er–0.08Zr alloy leads to an
thermally aged for 1 h. After each aging step, the samples were
water quenched to ambient temperature.
Vickers microhardness measurements were performed on
polished samples using a load of 200 gf and a dwell time of 10 s. At
least 10 independent indentations were made per sample across
different grains. The microstructural evolution was investigated by
transmission electron microscopy (TEM) using a JEOL 2100 with an
operating voltage of 200 kV. TEM foils were cut from the aged
specimens, subsequently ground to less than 100 μm and punched
into 3 mm diameter disks. Then the thin foils were prepared by
twin-jet polishing with an electrolyte solution consisting of 30%
nitric acid and 70% methanol at voltage  15 V DC and tempera-
ture below 25 °C. The size of precipitates and particle size dis-
tribution (PSD) of every condition was analyzed using Image-Pro
Plus software and by selecting more than 500 particles.

3. Results and discussion

Fig. 1 displays the age hardening behavior of Al–0.045Er– Fig. 1. Vickers microhardness evolution of Al–0.045E–0.08Zr [4], Al–0.045Er–
0.18Hf, and Al–0.045Er–0.08Zr–0.1Hf during isochronal aging as 0.18Hf and Al–0.045E–0.1Hf–0.08Zr alloys during isochronal aging.
determined by microhardness. Data for the ternary Al–0.04Er–
0.08Zr alloy [4] are plotted for comparison. In Al–0.045Er–0.18Hf
alloy, visible age hardening commences at 200 °C, and increases
slightly between 200 and 400 °C, then sharply increases at tem-
perature ranging from 400 to 475 °C as reflected by the variation
in the microhardness. The microhardness achieves a maximum
value of 609 MPa at 475 °C. Above 475 °C, the microhardness
continuously decrease as a result of the coarsening or dissolution
of the precipitates.
For Al–0.045Er–0.08Zr–0.1Hf alloy, the initial precipitation
hardening at temperature below 300 °C is similar to that of the Al–
0.045Er–0.18Hf and Al–0.04Er–0.08Zr alloys. However, the

Table 1
Compositions of the experimental alloys (at%).

Samples Nominal composition Verified composition

Er Zr Hf Er Zr Hf

Al–0.045Er–0.18Hf 0.045 – 0.18 0.045 – 0.185


Al–0.045Er–0.08Zr– 0.045 0.08 0.1 0.04 0.08 0.1
0.1Hf Fig. 2. Vickers microhardness evolution of Al–0.045E–0.08Zr [4], Al–0.045Er–
0.18Hf and Al–0.045E–0.1Hf–0.08Zr alloys during isothermal aging at 350 °C.
H. Wu et al. / Materials Science & Engineering A 639 (2015) 307–313 309

relationship:
D = D0 exp ( − Q /RT ) (1)

where D0 is the diffusion constant of the solute, R¼ 8.314 J/mol, K


is the universal gas constant, Q is the activation energy for diffu-
sion and T is the annealing temperature in K. The value of D0 and Q
are listed in Table 2, and then the diffusivity of Er [22], Zr [23,24]
and Hf [24] in Al at each given annealing temperature can be
calculated, as plotted in Fig. 4. It can be seen that the diffusivity of
solute in Al decreases from Er to Zr and Hf (i.e., DEr 4 DZr 4 DHf)
when the annealing temperature is lower than 567 °C. Therefore,
combining with the corresponding microstructures, the distinct
difference of precipitation hardening behaviors of experimental
alloys can be well explicated by the disparity of diffusivities of the
solute elements.
Fig. 5 shows the corresponding TEM micrographs of Al–
0.045Er–0.18Hf and Al–0.045Er–0.08Zr–0.1Hf after isochronal
aging. Only a few of precipitates can be detected in Al–0.045Er–
Fig. 3. Vickers microhardness evolution of Al–0.045Er–0.18Hf and Al–0.045E–
0.18Hf after isochronal aging to 400 °C and the corresponding
0.1Hf–0.08Zr alloys during isothermal aging at 375 °C.
hardness is only 400 MPa. The precipitates are coherent with the
Al matrix, as determined from the Ashby–Brown strain-field
increases of 100 MPa in hardness, but a delay of 20 h in the time to contrast, as shown in Fig. 5(a). But for the Al–0.045Er–0.08Zr–
obtain the peak hardness. Substituting 0.08Zr for Hf in an Al– 0.1Hf alloy after isochronal aging to the same temperature, much
0.045Er–0.18Hf leads to an increases of 20 MPa in hardness and an more precipitates can be detected as shown in Fig. 5(b), and the
advance of 20 h in the time to obtain the peak hardness. Addi- corresponding hardness is about 525 MPa which is much larger
tionally, the initial precipitation hardening rate of the Al–0.045Er– than that of the ternary Al–0.045Er–0.18Hf alloy. The two experi-
0.08Zr–0.1Hf, during aging time ranging from 1 h to 24 h, is sig- mental alloys have the same precipitation rate before isochronal
nificantly faster than that of the Al–0.045Er–0.18Hf. The hardness aging to 300 °C due to the precipitation of Er, because its diffu-
values of Al–0.045Er–0.18Hf and Al–0.045Er–0.08Zr–0.1Hf are sivity is considerably greater than the diffusivities of Zr and Hf. Zr
approximate equal (  300 MPa) between each other after aging for and Hf are essentially immobile in Al when aging at 300 °C for 3 h,
1 h, but the hardness values are 350 MPa and 450 MPa after aging with a root-mean-square (RMS) diffusion distance, 4Dt , of 0.52
for 24 h, respectively. However, the initial aging hardening rate of and 0.22 nm, respectively, as compared with RMS diffusion dis-
Al–0.04Er–0.08Zr alloy is similar to that of Al–0.045Er–0.08Zr– tance of 130.5 nm for Er. The corresponding RMS diffusion dis-
0.1Hf alloy, which could be attributed to the precipitation of Zr in tance of Zr and Hf at 400 °C is 22.8 nm and 9.6 nm, respectively. At
both of the above alloys. temperature ranging from 300 °C to 400 °C, the precipitation of Zr
Fig. 3 shows the precipitation hardening evolution of Al– accelerates the precipitation hardening rate, thus the precipitation
0.045Er–0.18Hf and Al–0.045E–0.1Hf–0.08Zr alloys during iso- rate of Al–0.045Er–0.08Zr–0.1Hf is larger than that of Al–0.045Er–
thermal aging at 375 °C. The peak hardness values are lower
compared with the peak values aged at 350 °C, especially for the Table 2
Reported diffusion data for Er, Zr and Hf in Al.
ternary alloy. The supersaturation of alloying elements decrease
with the increase of the aging temperature, then the volume Elements Er [23] Zr [24,25] Hf [25]
fraction of precipitation decreases correspondingly during aging at
1  12 2
a high temperature. It can also be seen that the precipitation
2
D0 (m s ) 4.3  10 7.28  10 1.07  10  2
Q (kJ mol  1) 77.2 242 241
evolution curves exhibit a great difference between Al–0.045Er–
0.18Hf and Al–0.045E–0.1Hf–0.08Zr alloys. The Al–Er–Hf alloy has
a long transient period with a slow increase in hardness values
after an incubation period, and then gets a peak hardness of
400 MPa. However, by substituting 0.08Zr for Hf in Al–0.045Er–
0.18Hf, the Al–0.045E–0.1Hf–0.08Zr alloy has a rapid precipitation
hardening rate and reach a peak hardness of 640 MPa which is
much higher than that of Al–0.045Er–0.18Hf alloy. It could be
attributed to the limited supersaturation of Hf in aluminum only at
375 °C, thus limits the volume fraction of precipitation. The solid
solubility of Zr in aluminum is close to zero at 375 °C [2] and Zr
can precipitate out almost completely. That is to say that the Al–
0.045E–0.1Hf–0.08Zr alloy have a larger supersaturation than that
of the Al–0.045Er–0.18Hf alloy. Therefore, Al–0.045E–0.1Hf–0.08Zr
alloy can obtain an significant strengthening effect after iso-
thermal aging at 375 °C.
Moreover, the precipitation kinetics is closely related to the
diffusivity of solute atoms in Al. Thus the distinct difference of
precipitation hardening behaviors of the three alloys should be
related to the consequence of sequential precipitation of solute
elements on the basis of the disparity in their intrinsic diffusivities. Fig. 4. Semi-logarithmic plot of diffusivity in Al vs. reciprocal temperature for the
The diffusion coefficient, D can be expressed by an Arrhenius-type elements of Hf, Zr and Er.
310 H. Wu et al. / Materials Science & Engineering A 639 (2015) 307–313

Fig. 5. TEM micrographs during isochronal aging: (a, c) bright-field image of Al–0.045Er–0.18Hf aged to 400 and 475 °C, respectively; (b, d) bright-field image of Al–0.045Er–
0.1Hf–0.8Zr aged to 400 and 450 °C, respectively; (e, d) EDS pattern of precipitates of Al–0.045Er–0.18Hf and Al–0.045Er–0.1Hf–0.8Zr, respectively.

0.15Hf during this period. Above temperature of 400 °C, due to the of the Al3Er, and the enveloping Zr/Hf-enriched shell as a barrier
precipitation of Hf, the two alloys have a similar precipitation can retard the growth of the precipitates. Since the diffusivity of Zr
hardening rate until the hardness reaches the peak value. is larger than that of Hf, then Zr segregate to the interfaces of pre-
A high number density of homogeneously distributed nearly existing Al3Er and slowing Al3Er growth at relatively low tem-
spherical precipitates can be seen in Fig. 5(c) and (d), which are perature than that of Hf, thus the mean radius of precipitates is
the micrographs of Al–0.045Er–0.18Hf and Al–0.045Er–0.08Zr– finer and the corresponding peak hardness is much larger in the
0.1Hf isochronal aged to the peak hardness. The volume fraction of quaternary alloy. The composition of the precipitates detected by
precipitates in two alloys are similar due to the similar total solute energy dispersive spectroscopy (EDS) consists of Al, Er, Hf in Al–
composition, but the average sizes are not equal. The inset at the 0.045Er–0.18Hf alloy and Al, Er, Hf, Zr in Al–0.045Er–0.08Zr–0.1Hf
lower right is the particle size distribution (PSD), it can be alloy, as shown in Fig. 5(e) and (f). The precipitates of the two
obtained that the average radius of the precipitates is about alloys have L12 structure as indicated by the superlattice diffrac-
3.4 nm in Al–0.045Er–0.18Hf alloy and only 2.5 nm in Al–0.045Er– tion point in the inserted selected area diffraction (SAD) pattern.
0.08Zr–0.1Hf, which may be attributed to that the pre-existing Therefore, the precipitates in the present Al–Er–Hf and Al–Er–Hf–
Al3Er precipitates grow up until Zr or Hf segregate to the interfaces Zr alloy should be L12 structured Al3(Er, Hf) and Al3(Er, Hf, Zr),
H. Wu et al. / Materials Science & Engineering A 639 (2015) 307–313 311

Fig. 7. Vickers microhardness vs. annealing curve of the as-rolled Al–0.045Er–


0.18Hf and Al–0.045Er–0.08Zr–0.1Hf alloys.

Fig. 8 shows the microstructure of the cold-rolled experimental


alloys after annealing for 1 h at 375 and 425 °C. When the alloys
were annealed at 375 °C, the recovery process started to emerge
and some newly formed subgrains were found along the rolling
direction in Al–0.045Er–0.18Hf alloy, as shown in Fig. 8(a). How-
ever, the microstructure of Al–0.045Er–0.1Hf–0.08Zr alloy is
composed of deformed grains containing a high density of tangled
dislocations (Fig. 8(b)). After annealing at 425 °C, partial recrys-
tallization started to emerge, and a small number of huge recrys-
tallized grains surrounded some subgrains can be seen in Al–
Fig. 6. TEM micrographs of alloys aged isochronally to peak hardness: (a) bright- 0.045Er–0.18Hf alloy, as shown in Fig. 8(c). Compared with the
field image of Al–0.045Er–0.18Hf; (b) dark-field image of Al–0.045Er–0.1Hf–0.8Zr. hardness of the as-rolled alloy, the corresponding hardness after
annealing at 425 °C has fallen by half, as indicated in Fig. 7. During
annealing at the same temperature of 425 °C, however, only the
respectively. Such similar structures of the precipitates and the
recovery process has emerged in the Al–0.045Er–0.1Hf–0.08Zr
synergy effects between joint added elements have been reported
alloy. The additional thermal energy permits the dislocations to
in published studies of Al–Er–Zr [4,11], Al–Sc–Hf [18], Al–Sc–Er–Zr
move and form the boundaries of a polygonized subgrain structure
[26], Al–Sc–Hf–Zr [18] alloys, etc.
as indicated in Fig. 8(d). Compared with Al–0.045Er–0.18Hf alloy,
Fig. 6(a) and (b) is the bright-field image of Al–0.045Er–0.18Hf
Al–0.045Er–0.1Hf–0.08Zr alloy possess a more outstanding
and dark-field image of Al–0.045Er–0.08Zr–0.1Hf alloys after iso-
recrystallization resistance, which can be related to the different
thermal aging at 350 °C for 100 and 84 h, respectively. And both of
number density of precipitates, i.e. the f/r ratio in the two alloys.
them are corresponding to the precipitates in peak aging state. The
The normal process of recrystallization in metals consists of
dense dispersed precipitates are homogeneously distributed, and
formation of strain-free nuclei/sub-grains and the subsequent
the inset SAD pattern implies that the precipitates have the L12 growth of these into the surrounding deformed material. The
structured lattice, as shown in Fig. 6(a). The volume fraction of critical size (RC) of a nucleus can be described mathematically by
precipitates in two alloys are approximately equal, but the average the following Gibbs–Thompson relationship [1]:
diameters of precipitates are approximately 2.3 and 1.9 nm for Al–
0.045Er–0.18Hf and Al–0.045Er–0.08Zr–0.1Hf alloys, respectively, 4γGB
R > RC =
which can been obtain from the PSD illustrations at lower right of PD − PZ (2)
Fig. 6(a) and (b), so the peak hardness of Al–0.045Er–0.08Zr–0.1Hf
where RC is the critical radius for nucleation, γGB is the specific
is slight larger than that of Al–0.045Er–0.18Hf.
grain boundary energy and PD is the stored deformation energy. PZ
is the retarding force (usually described as the Zener drag) that the
3.1. Recrystallization behavior dispersoids exert upon the dislocation and subgrain boundaries. A
widely used estimate for the Zener darg is [1]
The hardness as a function of annealing temperature curves of
3fγGB
the as-rolled peak aged Al–0.045Er–0.18Hf and Al–0.045Er–0.1Hf– PZ =
2r (3)
0.08Zr alloys is given in Fig. 7. The recrystallization temperature
was defined as the temperature at which the hardness value where r is the radius and f is the volume fraction of dispersoids.
decreases to 50%. The microhardness of Al–0.045Er–0.18Hf According to Eqs. (2) and (3), in order to achieve a high recrystalli-
decreases to 377 MPa at 425 °C, while Al–0.045Er–0.1Hf–0.08Zr zation resistance, a high volume fraction of dispersoids (high f/r
maintains a relatively high microhardness of 561 MPa until 450 °C. ratio) is necessary, or in other words a high volume fraction of fine
The recrystallization temperature of Al–0.045Er–0.1 Hf–0.08Zr is particles result in the most effective retardation of dislocation and
450 °C, about 25 °C higher than the recrystallization temperature boundary migration. The number densities of precipitates of both
of 425 °C for the Al–0.045Er–0.18Hf alloy. Al–0.045Er–0.1Hf–0.08Zr and Al–0.045Er–0.18Hf alloys peak aged
312 H. Wu et al. / Materials Science & Engineering A 639 (2015) 307–313

Fig. 8. Bright-field TEM images of as-rolled alloys annealed at different temperatures for 1 h: (a, c) Al–0.045Er–0.18Hf alloy annealing at 375 and 425 °C, respectively; (b, d)
Al–0.045Er–0.1Hf–0.08Zr alloy annealing at 375 and 425 °C, respectively.

isothermally at 350 °C are similar to each other. Due to the average precipitates in both alloys and its effect on recrystallization resis-
radius of precipitates in the Al–0.045Er–0.1Hf–0.08Zr is finer than tance will be carried out in our future work.
that of Al–0.045Er–0.18Hf, as indicated in Fig. 5, thus the as-rolled
Al–0.045Er–0.1Hf–0.08Zr alloy with a higher f/r ratio should have a
higher recrystallization temperature.
Fig. 9(a, c) and (b, d) shows the fully recrystallized micro- 4. Conclusion
structures for the Al–0.045Er–0.18Hf and Al–0.045Er–0.1Hf–0.08Zr
alloys after annealing at 500 °C for 1 h, respectively. The high- The precipitation hardening and recrystallization behavior of
angle boundaries of the recrystallized grains can be seen in Fig. 9 Al–Er–Hf–Zr alloys have been studied by microhardness mea-
(a, b). The precipitates are found to be significantly coarse in surements and transmission electron microscopy observation. The
comparison with that of the as-rolled alloys. Dislocations and results show that the two experimental alloys exhibit an excellent
grain/sub-grain boundaries provide the pipeline diffusion chan- precipitation hardening effect and recrystallization resistance,
nels, through which diffusion of solutes atoms is much easier than especially the Al–0.045Er–0.1Hf–0.08Zr alloy. The peak hardness
that in the matrix. As a result, precipitates will coarsen rapidly in values of the Al–0.045Er–0.08Zr–0.1Hf are 644 MPa and 662 MPa
the rolled alloys especially with increasing annealing temperature. after isochronal aging to 450 °C and isothermal aging at 350 °C for
The average size of the precipitates increases and the number 80 h, respectively. Due to the sequential precipitation of solute
density decreases compared with that of as-rolled alloy, then the
elements on the basis of the disparity in their intrinsic diffusivities
crystallization resistance decreases. However, it can also be seen
(DEr 4DZr 4DHf), substituting 0.08Zr for Hf in an Al–0.045Er–
that the coarse precipitates seemed to act as strong nails causing a
0.18Hf can accelerate the precipitation hardening rate to achieve
severe boundary distortion, which indicates that hindrance by
the peak hardness. The average size of precipitates at peak aged
precipitates still plays an important role on the growth of the
state in Al–0.045Er–0.1Hf–0.08Zr is much finer than that of the Al–
recrystallized grains, even when the alloys are fully recrystallized.
0.045Er–0.18Hf alloy. The recrystallization temperature of Al–
Fig. 9(c, d) displays the precipitates in the recrystallized grain
interior of Al–0.045Er–0.18Hf and Al–0.045Er–0.1Hf–0.08Zr alloy, 0.045Er–0.1Hf–0.08Zr is 450 °C, about 25 °C higher than that of the
respectively. It can be found that the average size of the pre- Al–0.045Er–0.18Hf (425 °C). The high recrystallization resistance
cipitates in quaternary alloy is about 35 nm, significantly smaller can be attributed to the larger f/r ratio of precipitates, which leads
than that of the ternary alloy (153 nm). The difference in the to a higher Zener drag (PZ) to resist recrystallization in Al–
growth rate of the particle size also must have influence on the 0.045Er–0.08Zr–0.1Hf alloy. The precipitation in Al–0.045Er–
anti-recrystallization performance of these alloys. Therefore, the 0.1Hf–0.08Zr alloy is more stable at elevated temperature than
thermal stability (coarsening resistance and phase stability) of the that in the Al–0.045Er–0.18Hf alloy.
H. Wu et al. / Materials Science & Engineering A 639 (2015) 307–313 313

Fig. 9. Bright-field TEM images of as-rolled alloys annealed at 500 °C for 1 h: (a, c) Al–0.045Er–0.18Hf alloy; (b, d) Al–0.045Er–0.1Hf–0.08Zr alloy.

Acknowledgments [10] B. Forbord, W. Lefebvre, F. Danoix, H. Hallem, K. Marthinsen, Scr. Mater. 51


(2004) 333–337.
[11] H.Y. Li, J. Bin, J.J. Liu, Z.H. Gao, X.C. Lu, Scr. Mater. 67 (2012) 73–76.
This work was supported in part by the National Key Basic [12] S.P. Wen, K.Y. Gao, Y. Li, H. Huang, Z.R. Nie, Metall. Mater. Trans. A 44A (2013)
Research & Development Plan Project (No. 2012CB619503), 2849–2856.
National and Beijing Natural Science Foundation Project (Nos. [13] Y. Zhang, W. Zhou, H. Gao, Y. Han, K. Wang, Scr. Mater. 69 (2013) 477–480.
[14] Z.H. Gao, H.Y. Li, J.J. Liu, X.C. Lu, Y.X. Ou, J. Alloy. Compd. 592 (2014) 100–104.
51201003 and 2142007), the National High Technology Research [15] S.P. Wen, K.Y. Gao, Y. Li, H. Huang, Z.R. Nie, J. Alloy. Compd. 599 (2014) 65–70.
and Development Program (No. 2013AA031301), International [16] H.Y. Li, Z.H. Gao, H. Yin, H.F. Jiang, X.J. Su, J. Bin, Scr. Mater. 68 (2013) 59–62.
Science & Technology Cooperation Program of China (No. [17] V.V. Zakharov, T.D. Rostova, Met. Sci. Heat Treat. 49 (2007) 435–442.
[18] H. Hallem, W. Lefebvre, B. Forbord, F. Danoix, K. Marthinsen, Mater. Sci. Eng. A
2013DFB50170). 421 (2006) 154–160.
[19] E. Clouet, L. Lae, T. Epicier, W. Lefebvre, M. Nastar, A. Deschamps, Nat. Mater.
(2006) 482–488.
[20] W. Lefebvre, F. Danoix, H. Hallem, B. Forbord, A. Bostel, K. Marthinsen, J. Alloy.
References Compd. 470 (2009) 107–110.
[21] K.E. Knipling, R.A. Karnesky, C.P. Lee, D.C. Dunand, D.N. Seidman, Acta Mater.
[1] E. Nes, N. Ryum, O. Hunderi, Acta Metall. 33 (1985) 11–22. 58 (2010) 5184–5195.
[2] K.E. Knipling, D.C. Dunand, D.N. Seidman, Z. Metallkunde 97 (2006) 246–265. [22] H. Hallem (Ph.D. thesis), Norwegian University of Science and Technology,
[3] K.E. Knipling, D.C. Dunand, D.N. Seidman, Metall. Mater. Trans. A 38 (2007) 2005.
2552–2563. [23] Y. Zhang, K.Y. Gao, S.P. Wen, H. Huang, Z.R. Nie, D.J. Zhou, J. Alloy. Compd. 610
[4] S.P. Wen, K.Y. Gao, Y. Li, H. Huang, Z.R. Nie, Scr. Mater. 65 (2011) 592–595. (2014) 27–34.
[5] K.E. Knipling, D.N. Seidman, D.C. Dunand, Acta Mater. 59 (2011) 943–954. [24] K. Hirano, S.I. Fujikawa, J. Nucl. Mater. 69 (1978) 564–566.
[6] H. Hallem, B. Forbord, K. Marthinsen, Mater. Sci. Eng. A 387–389 (2004) [25] S.I. Fujikawa, J. Jpn. Inst. Light Met. 46 (1996) 202–215.
940–943. [26] C. Booth-Morrison, D.C. Dunand, D.N. Seidman, Acta Mater. 59 (2011)
[7] S.K. Pandey, C. Suryanarayana, Mater. Sci. Eng. A 111 (1989) 181–187. 7029–7042.
[8] H. Wu, S.P. Wen, K.Y. Gao, H. Huang, W. Wang, Z.R. Nie, Scr. Mater. 87 (2014)
5–8.
[9] M.E. Van Dalen, R.A. Karnesky, J.R. Cabotaje, D.C. Dunand, D.N. Seidman, Acta
Mater. 57 (2009) 4081–4089.

You might also like