You are on page 1of 4

Available online at www.sciencedirect.

com

Scripta Materialia 66 (2012) 911–914


www.elsevier.com/locate/scriptamat

High-temperature deformation and structural restoration


of a nanostructured Al alloy
H. Asgharzadeh,a A. Simchib,c,⇑ and H.S. Kimd
a
Department of Materials Engineering, Faculty of Mechanical Engineering, University of Tabriz, P.O. Box 51666-16471, Tabriz, Iran
b
Department of Materials Science and Engineering, Sharif University of Technology, P.O. Box 11365-9466, Tehran, Iran
c
Institute for Nanoscience and Nanotechnology, Sharif University of Technology, P.O. Box 11365-9466, Tehran, Iran
d
Department of Materials Science and Engineering, Pohang University of Science and Technology, P.O. Box 790-784, Pohang,
South Korea
Received 14 November 2011; revised 9 February 2012; accepted 13 February 2012
Available online 24 February 2012

We studied the flow stress and microstructural changes of nanostructured Al-6063 alloy produced by mechanical alloying at var-
ious temperatures and strain rates. The analysis of flow curves was performed by a constitutive equation, and the stress exponent
and activation energy were determined as functions of strain. The deformation mechanisms were elaborated through microstructural
observations by electron backscattering diffraction and transmission electron microscopy. Coarsening of the subgrains and grain
growth upon deformation was monitored and related to the Zener–Hollomon parameter.
Ó 2012 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Aluminum; Nanostructured materials; High-temperature deformation; Restoration mechanisms

Interest in the field of nanostructured (NS) mate- temperatures may be controlled by dislocation climb
rials has grown dramatically in the last decade due to as well [9].
their unique microstructures and mechanical perfor- In general, NS materials have a large area of grain
mance [1]. Bulk NS metals are frequently produced by boundaries and therefore a large stored energy, and are
severe plastic deformation (SPD) techniques, such as intrinsically unstable with respect to grain growth during
equal channel angular extrusion [2], high-pressure the high-temperature deformation operation [10]. There-
torsion [3] and mechanical milling followed by hot fore, it is imperative and useful to study the deformation
consolidation methods [4]. NS materials exhibit higher behavior of NS materials and to determine the restora-
strength and lower ductility and formability compared tion mechanisms dependent on temperature, strain and
to their coarse-grained (CG) counterparts [5]. In strain rate. As a follow-up to our recent work on
spite of significant studies on the room-temperature room-temperature mechanical behavior of NS-Al6063
properties of NS materials, their mechanical properties alloy [11], we herein report the microstructural evolution
at elevated temperatures, particularly restoration mech- of the alloy during high deformation at different temper-
anisms involved in microstructural changes occurring atures and strain rates. The effects of thermomechanical
during the deformation, have been reported only rarely. parameters on the flow behavior and dynamic restora-
Horita et al. [6] and Park et al. [7] showed that the fine tion mechanisms are presented and discussed.
grain structure of SPD-processed Al alloys could be Details of the processing method of the NS Al alloy can
stable up to 300 °C (0.5Tm), whereas consistent grain be found in Ref. [12]. Briefly, gas-atomized Al-6063 pow-
growth occurred at higher temperatures. Ma et al. [8] der, with a chemical composition of Al–0.64Mg–0.67Si–
observed superplastic behavior in an ultrafine-grained 0.32Fe–0.2Cu (in wt.%), was utilized. Mechanical alloy-
Al alloy through grain boundary sliding. The deforma- ing was performed in an attrition ball mill (Union Process,
tion mechanism of NS aluminum alloys at elevated OH, USA) for 20 h under an Ar atmosphere. Stearic acid
(1.5 wt.%) was added as process control agent (PCA). The
powder was degassed at 400 °C, compacted in an alumi-
⇑ Corresponding author at: Department of Materials Science and num container and finally extruded at 450 °C with an
Engineering, Sharif University of Technology, P.O. Box 11365-9466, extrusion ratio of 14:1. Cylindrical specimens with diam-
Tehran, Iran. Tel./fax: +98 21 6616 5262; e-mail: simchi@sharif.edu eter of 4 mm and an aspect ratio of 1.2 were machined

1359-6462/$ - see front matter Ó 2012 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.scriptamat.2012.02.026
912 H. Asgharzadeh et al. / Scripta Materialia 66 (2012) 911–914

from the extruded billet. Uniaxial compression test was g). Consequently, a high value of the average misorien-
performed on a Caster and Thermomechanical Simulator tation angle was measured (have  33°). Figure 2j
(Fuji Electronic Industrial Co., Ltd, Japan) under a high- illustrates a representative TEM image of the micro-
purity Ar atmosphere in a temperature of 300–450 °C and structure. A high density of dislocations is observable
a strain rate range of 0.01–1 s1. Quartz plates were used within larger grains while nanometric precipitates are
for reducing the friction between the specimen and anvils. distributed in the microstructure. Figure 2b and c shows
A soaking time of 3 min at the testing temperature was that a significant change in the grain structure occurred
considered for temperature stabilization. To preserve after hot deformation. The elongated grains were mainly
the microstructure after deformation, the specimens were disappeared and nanostructured and ultrafine equiaxed
instantly quenched with a cooling rate of 70 °C s1. grains were formed after deformation at 300 °C and
The microstructural studies before and after defor- 1 s1 (Fig. 2b). The size distribution of grains became
mation were carried out by electron backscattering dif- narrower but the average grain size was slightly in-
fraction (EBSD; Hikari, EDAX, NJ), using a field creased (Fig. 2e). Moreover, the values of fHAB and have
emission gun scanning electron microscope (Helios were declined (Fig. 2h). To reveal the effect of tempera-
NanoLab DualBeam, FEI, Oregon), and transmission ture, the microstructure of the specimen deformed at
electron microscopy (TEM; JEM-2100F STEM, Japan). 450 °C and 1 s1 is shown in Figure 2c. Micrometric
Figure 1 shows true stress–true strain curves of NS substructured grains and ultrafine equiaxed grains are
Al-6063 at various temperatures and strain rates. The seen. Moreover, the grain size distribution is broader
true stress was corrected for friction according to the (Fig. 2f) and the average grain size is coarser (760 nm).
slab analysis of the compression of a cylinder [13]. The Furthermore, a shift in the boundary misorientation dis-
flow behavior dependent on the temperature and strain tribution from high angles to low angles is noticed
rate is similar to that of CG aluminum alloy, as reported (Fig. 2i). As a result, the values of fHAB and have changed
elsewhere [14]. As seen, the flow stress expeditiously in- to 0.67 and 27°, respectively. Meanwhile, the stability of
creases to a maximum value and then gradually de- precipitates during high-temperature deformation is
creases with the true strain. This behavior is due to a noticeable (Fig. 2k and l). The microstructural observa-
dynamic competition between the work hardening tions determine that DRV through the formation of
caused by interaction, pile-up and tangling of disloca- subgrains within the grains and DRX via formation of
tions, and the work softening may be caused by dynamic recrystallized grains occurred during hot deformation
recovery (DRV), subgrain coarsening, dynamic precipi- of NS Al-6063.
tation, dynamic recrystallization (DRX) and texture In order to determine the restoration mechanisms
softening [15]. involving in the deformation behavior, the following
The microstructural features of the alloy before and hyperbolic sine law between the flow stress (r), temper-
after hot deformation at some selected conditions are ature (T) and strain rate (_e) was utilized [16]:
shown in Figure 2. In EBSD inverse pole-figure (IPF)  
n Q
maps, low-angle boundaries (LABs) (misorientation be- A½sinhðarÞ ¼ e_ exp ¼Z ð1Þ
tween adjacent grains of 2°–15°) and high-angle bound- RT
aries (HABs) (misorientation larger than 15°) are shown where Q is the deformation activation energy, R is the
with white and black lines, respectively. The IPF image gas constant, n is the stress exponent, Z is the Zener–
of the as-extruded material reveals an elongated grain Hollomon parameter, and A and a are material con-
structure with an aspect ratio of 2.9 (Fig. 2a). As shown stants. The stress multiplier for the highest correlation
in Figure 2d, the microstructure consists of grains with a coefficient for the linear correlation between ln e_ and
wide size distribution in the ultrafine size range (>95%) ln sinh(ar) was determined to be a = 0.06 MPa1, which
and an average value of 500 nm. While the microstruc- is close to the value obtained for the CG Al alloys [16].
ture consisted of a high fraction of HABs (fHAB  0.78), The values of n and Q at different strains (e) were deter-
some of the larger grains contain subgrains (Fig. 2a and mined from the slope of ln e_ vs. ln sinh(ar) plot and Rn
ln sinh(ar) vs. 1/T, respectively. An example of the pro-
cedure is shown in Supplementary Figure 3a and b for
e = 0.5. Figure 3 depicts the variation of activation en-
ergy and stress exponent with the true strain. Such
dependency has also been observed for a number of
CG Al alloys, such as AlFe8.5V1.3Si1.7 [17]. Data fitting
yields:
n ¼ 16:7e5  108 e4 þ 9:2e3 þ 3e2  4:7e þ 3:8 ð2Þ

Q ¼ 2441:7e5 þ 8:3e4 þ 1737:1e3  232:6e2  323:3e þ 298:4


ð3Þ
Although these equations are curve fitting expres-
sions and might not have a physical basis, an average
value of 2.9 ± 0.2 and 228 ± 21 kJ mol1 could be con-
Figure 1. True stress–strain curves of NS Al-6063 alloy at various sidered for n and Q, respectively, considering that the
temperatures and strain rates. variation of Q reflects changes in n, as explained else-
H. Asgharzadeh et al. / Scripta Materialia 66 (2012) 911–914 913

Figure 2. Microstructural analysis of NS-Al6063 alloy before (left-hand-side images) and after deformation at 300 °C and 1 s1 (central images) and
450 °C and 1 s1 (right-hand-side images): (a–c) orientation maps from EBSD measurement (the insert in (a) shows a representation of the color code
used to identify the crystallographic orientations on standard stereographic projection); (d–f) grain size distribution from EBSD measurement; (g–i)
misorientation angle distribution from EBSD measurement; (j–l) TEM micrographs.

deformation. The higher activation energy can also be


linked to the occurrence of DRX. The progressive accu-
mulation of dislocations in LABs increases their misori-
entation, leading to the transformation of LABs into
HABs. As the straining progresses, the subgrains, grains
and precipitates coalesce, easing dislocation motion so
that the activation energy falls (Fig. 3).
Figure 4a shows variation in the size of subgrains (dS)
and recrystallized grains (dR) as functions of ln Z.
Although lower values for dS and dR were obtained com-
pared to conventional Al alloys [18–20], these are com-
parable with those reported for SPD-processed Al
Figure 3. Constitutive analysis of NS Al-6063 in hot compression.
alloys with similar Z parameters [21]. Similar to CG
Stress exponent (n) and activation energy (Q) as functions of true Al alloys [22], linear relationships in semi-logarithmic
strain (symbols and dashed lines show experimental data and the scale can be established between the grain structure
polynomial fitted curves, respectively). and Z, i.e. d = a  b ln Z, where a and b are the fitting
parameters. The correlation of the dS and dR with the
steady-state flow stress (rss) in logarithmic scale is illus-
where [14]. Meanwhile, values in the range of 155– trated in Figure 4b. Similar to conventional alloys [23],
205 kJ mol1 for the activation energy of CG Al–Mg– the graph shows that the relationship between the size
Si alloys of various purities at pre-aged or annealed con- of subgrains and recrystallized grains with rss during
ditions have been reported [16]. The higher activation hot working can be expressed as rss = Kdm. However,
energy for the nanostructured Al-6063 can be attributed analysis of the data yielded m values of 2.56 and 1.04
to the microstructural features (nanostructured grains, for subgrains and recrystallized grains, respectively,
ultrafine grains and subgrains) that limit the cross slip which are higher than those reported for the hot defor-
and climb processes, especially at the initial stages of mation of micro-crystalline alloys [24,25].
914 H. Asgharzadeh et al. / Scripta Materialia 66 (2012) 911–914

curve. An apparent activation energy of


228 ± 21 kJ mol1 was obtained. The coarsening of the
subgrains, recrystallized grains and precipitates during
hot deformation, especially at low Z values, was shown.
Deformation at Z > 3.4  1017 s1 resulted in a refine-
ment of grain structure without a significant coarsening
of subgrains.

Supplementary data associated with this article can


be found, in the online version, at doi:10.1016/
j.scriptamat.2012.02.026.

[1] V.L. Tellkamp, A. Melmed, E.J. Lavernia, Metall. Mater.


Trans. 32A (2001) 2335.
[2] R.Z. Valiev, T.G. Langdon, Prog. Mater. Sci. 51 (2006)
881.
[3] A.P. Zhilyaev, T.G. Langdon, Prog. Mater. Sci. 53 (2008)
893.
[4] C. Suryanarayana, Prog. Mater. Sci. 46 (2001) 1.
[5] Y.M. Wang, E. Ma, Acta Mater. 52 (2004) 1699.
[6] Z. Horita, T. Fujinami, M. Nemoto, T.G. Langdon,
Metall. Mater. Trans. 31A (2000) 691.
[7] K.-T. Park, H.-J. Kwon, W.-J. Kim, Y.-S. Kim, Mater.
Sci. Eng. A316 (2001) 145.
[8] Z.Y. Ma, F.C. Liu, R.S. Mishra, Acta Mater. 58 (2010)
4693.
[9] X.L. Shi, R.S. Mishra, T.J. Watson, Scripta Mater. 52
(2005) 887.
Figure 4. Dependency of the size of subgrains and recrystallized grains [10] F.J. Humphreys, P.B. Prangnell, R. Priestner, Curr. Opin.
on (a) the Zener–Hollomon parameter and (b) the steady-state flow Solid State Mater. Sci. 5 (2001) 15.
stress. [11] H. Asgharzadeh, A. Simchi, H.S. Kim, Metall. Mater.
Trans. 42A (2011) 816.
[12] H. Asgharzadeh, A. Simchi, H.S. Kim, Mater. Sci. Eng. A
The results indicate that significant changes in the
528 (2011) 3981.
microstructure of the NS alloy occur during the hot [13] P. Cavaliere, Composites 35A (2004) 619.
deformation, although fine particles are distributed in [14] H. Asgharzadeh, A. Simchi, H.S. Kim, Master. Sci. Eng.
the matrix (Fig. 2). As reported elsewhere [12], these A (2012), doi:10.1016/j.msea.2012.02.031.
particles are mainly Al8Fe2Si, Al5FeSi and Al8Mg3Fe- [15] B. Verlinden, A. Suhadi, L. Delaey, Scripta Metall.
Si2, with a total of 1.8 vol.%. According to the theory Mater. 28 (1993) 1441.
of Zener pinning, the volume fraction (Vf) of particles [16] H.J. McQueen, N.D. Ryan, Mater. Sci. Eng. A 322 (2002)
with a diameter of dP required to stabilize a microstruc- 43.
ture with an average grain size of D should be Vf = 4dp/ [17] M.Y. Zhan, Z. Chen, H. Zhang, W. Xia, Mech. Res.
3D [1]. By considering dP = 62 nm [12], the required Vf Commun. 33 (2006) 508.
[18] Y. Deng, Z. Yin, J. Huang, Mater. Sci. Eng. A 528 (2011)
to stabilize the nanosize grains was estimated to be
1780.
16.5%. Therefore, effective grain boundary pinning can- [19] H. Jazaeri, F.J. Humphreys, Acta Mater. 52 (2004) 3251.
not be afforded by the particulates. It is pertinent to [20] T. Furu, H.R. Shercliff, G.J. Baxter, C.M. Sellars, Acta
point out that, upon mechanical milling and in the pres- Mater. 47 (1999) 2377.
ence of PCA, carbides and oxides can be formed and [21] O. Sitdikov, T. Sakai, E. Avtokratova, R. Kaibyshev, Y.
distributed within the Al matrix, and will contribute to Kimura, K. Tsuzaki, Mater. Sci. Eng. A 444 (2007) 18.
the dispersion strengthening [4,12]. We evaluated this ef- [22] C. Poletti, M. Rodriguez-Hortalá, M. Hauser, C. Som-
fect by analyzing the threshold stress requiring for dislo- mitsch, Mater. Sci. Eng. A 528 (2011) 2423.
cation motion at elevated temperatures [26] for NS Al- [23] T. Furu, K. Marthinsen, E. Nes, Recrystallization’92,
6063 and obtained an approximate value of 1 MPa. This Spain, 1992, p. 41.
[24] F.J. Humphreys, M. Hatherly, Recrystallization and
observation reflects the minor influence of these parti-
Related Annealing Phenomena, second ed., Elsevier,
cles, which have on the hot deformation behavior. Oxford, 2004.
In conclusion, we studied the deformation behavior [25] H.J. McQueen, C.A.C. Imbert, J. Alloy Compd. 378
and microstructural changes of a nanostructured Al- (2004) 35.
6063 alloy across a wide range of temperatures and [26] S. Spigarelli, E. Evangelista, H.J. McQueen, Scripta
strain rates. The stress exponent and activation energy Mater. 49 (2003) 179.
were correlated to strain with a fifth-order polynomial

You might also like