Vous êtes sur la page 1sur 4

Available online at www.sciencedirect.

com

Scripta Materialia 68 (2013) 861864


www.elsevier.com/locate/scriptamat

Size eect on the deformation behavior of duralumin micropillars


R. Gu and A.H.W. Ngan
Department of Mechanical Engineering, The University of Hong Kong, Pokfulam Road, Hong Kong, Peoples Republic of China
Received 26 January 2013; revised 5 February 2013; accepted 6 February 2013
Available online 19 February 2013

Pure metal microspecimens have been found to exhibit strong size dependence of strength, but alloyed counterparts with a much
rened microstructural length scale due to the precipitates present are as yet unknown. Here, compression tests on duralumin (aluminum 2025 alloy) micropillars reveal a much weaker size dependence of strength compared to pure Al, indicating the predominance
of the internal length scale in determining strength. Creep is also signicant in duralumin, probably due to the viscous overcoming of
obstacles during deformation.
2013 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Keywords: Plastic deformation; Compression tests; Dislocations; Nanoindentation; Precipitation hardening

The strength of micron-sized metallic single crystals has been found to depend on their size generally in
accordance with a power law [13], with jerky stress
strain behavior [1,48]. Such unusual deformation behavior is now explainable by a number of mechanisms which
have received direct experimental support [8,9]. The dislocation starvation [6] theory considers that small crystals can remain in a dislocation-free state as the mobile
dislocations glide and are eliminated at free surfaces without accumulation and multiplication. The source truncation [10] and exhaustion hardening [11] models
consider the distribution and operation of dislocation
sources in a small conned volume, leading to the breakdown of a mean-eld condition for forest hardening. The
power law dependence of the strength on size has also
been explicitly predicted as a consequence of Taylor-type
hardening in an initial fractal dislocation network [12].
While the jerky deformation of monolithic microcrystals is due to the ease of loss of dislocations from the
specimen volume, several studies have investigated
methods to trap dislocations inside the specimen, such
as coating its surface [13,14]. Experiments on microsized
bicrystals have also shown improved strengthening and
suppressed serrated ow when compared with the single
crystalline state [15], but in nanosized bicrystals, the
grain boundary could act as a dislocation sink, leading
to adverse eects [16]. The deformation behavior of micro- and nanospecimens with polycrystalline [17] and
even nanocrystalline [1820] substructures has also been

Corresponding author. Tel.: +852 2859 2624; e-mail: gurui@hku.hk

found to be dependent on the ratio between the grain


size and the specimen size, and nanospecimens containing nanoscale twins have also attracted some attention
[2022] due to their high strength and ductility.
In this study, we investigate another way to change
the internal microstructural length scale in a small crystal, namely, by introducing second-phase precipitates, to
block dislocations. The deformation behavior of micropillars fabricated from a precipitate-hardened aluminum
alloy was studied by controlling the internal length scale
of the precipitate microstructure by heat treatment.
A bulk piece of aluminum alloy 2025 (Al4 wt.% Cu
1.3 wt.% Mg1.3 wt.% Ag0.6 wt.% Mn) was cut into
several pieces, which were then homogenized at 500 C
for 24 h. The specimens were solution heat-treated at
520 C for 3 h in an air furnace, followed immediately
by a water quench in icy water. They were then stretched
by 25% in a tensile machine, in order to reduce the vacancies in the specimen, which would be an adverse condition
for precipitate growth in the following step. The bulk
specimens were then aged at 170 C for 4, 10, 15, 20 and
25 h, respectively. Their Vickers hardness was measured
and the peak-aged condition for the highest hardness of
125 Hv was found to be at 15 h of aging. In addition
to this peak-aged state, another condition, without articial aging but naturally aged at room temperature (RT)
for dozens of days, with a hardness 110 Hv, was chosen
for the subsequent experiments. The specimens from both
states were treated by mechanical polishing, then electropolishing in a 1:4 mixture of nitric acid and methanol at
30 C and 15 V for 30 s. Afterwards, on a bulk sample
of each of the two aged states, a large grain with an

1359-6462/$ - see front matter 2013 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.scriptamat.2013.02.012

862

R. Gu, A. H. W. Ngan / Scripta Materialia 68 (2013) 861864

orientation of [0 1 1] and a diameter larger than 100 lm


was selected for the fabrication of micropillars, by focused
ion-beam (FIB) milling in a Quanta 200 3D Dual Beam
FIB/SEM system operating at an ion beam voltage of
30 kV, using a series of concentric annular-pattern millings with varied currents. Micropillars of diameters from
6.5 to 1.0 lm and with a diameter-to-height ratio of
1:3.5, were fabricated. Compression of the micropillars
was performed at room temperature in an Agilent G200
Nanoindenter with a at-ended diamond punch in a
load-controlled manner, with a constant loading and
unloading rate of 12 MPa s1. For dierent pillar sizes,
the resultant strain rate was on the order of 103 s1.
Scanning electron microscopy (SEM) examination was
carried out in a LEO1530 microscope, and transmission
electron microscopy (TEM) was performed in a Philips
Tecnai microscope operating at 200 kV. The TEM samples were fabricated from 10% compressed pillars by
longitudinal sectioning by FIB milling using a protocol
previously reported [14].
Figure 1(a and b) show the stressstrain curves of RTaged and peak-aged duralumin pillars along the [0 1 1] orientation. In the initial loading stage, the strength in both
aged states increases as the pillar size gets smaller. Figure 2
shows that the 2% proof stress and the 10% ow stress of
the pillars decrease with diameter D approximately in a
power law ry / Dm . For the 2% proof strength, the exponent m is 0.37 for the RT-aged state and 0.51 for the peakaged state; for the 10% ow stress, the size eect is even
weaker, with m decreasing to 0.07 and 0.15 for RT-aged
and peak-aged states, respectively. The much lower m values at 10% strain mean that the size eect almost disappears at high strain levels, especially for the RT-aged
condition.
Figure 1 also shows that the 1 lm duralumin pillars
deform in a jerky manner, with strain bursts often larger
than 0.2%. However, when the pillar size is increased,
the stressstrain curves are much smoother. Figure 3(a
and b) shows typical SEM images of the deformed
duralumin micropillars aged at peak condition. Interestingly, close inspection of Figure 3(a) reveals that the
5 lm duralumin pillar here has only one type of dense
and ne slip step up to 0.02 lm in size, and these are
uniformly distributed on the pillars surface. Since the
strains associated with these slip steps are at most
0.1%, it is not surprising that the stressstrain curve
of this pillar size in Figure 1(a) appears to be smooth.
The slip steps on the 1 lm duralumin pillar in Figure 3(b)
are more inhomogeneous, with spacing up to 0.3 lm,
which correspond well to the larger strain bursts in the
stressstrain curve in Figure 1(a). Nevertheless, when

Figure 1. Stressstrain curves of micropillars of dierent sizes fabricated from an [0 1 1] grain on (a) RT-aged and (b) peak-aged
duralumin alloy, respectively.

Figure 2. 2% proof strength and 10% ow stress of duralumin


micropillars vs. pillar diameter.

Figure 3. SEM images of deformed peak-aged duralumin pillars of (a)


5 lm and (b) 1 lm diameter. The images were taken with 30 titling.

compared to pure Al pillars of the same size [7], the slip


steps in the duralumin pillars are in general smaller and
more homogeneously distributed.
Figure 1(a and b) also shows that a nose often appears at the onset of unloading. However, the unloading
curves in pure aluminum pillars [7,12,14] are quite linear, with the slope almost equal to the Youngs modulus
of aluminum of 70 GPa. The non-linear unloading
behavior in the duralumin pillars here is likely an eect
of creep of the specimen [23]. In the unloading stage,
assuming that strain is the sum of an elastic and a
time-dependent component, the strain rate is
e_ t e_ e t e_ t t, where the elastic strain rate is
_
rt
12MPas1

1:7  104 s1


e_ e t
E
70GPa
with an unloading rate of 12 MPa s1 being used, and
e_ t t is the time-dependent strain rate comprising the
specimens creep strain rate e_ c t and the machines thermal drift rate e_ th t, i.e. e_ t t e_ c t e_ th t. The machine
drift rate during compression tests was typically
0.5 nm s1 and, taking a typical peak-aged pillar size
of 6.5 lm as an example, e_ th t should be
2.5  105 s1 based on the pillars height with the
1:3.5 aspect ratio. The overall strain rate e_ t during
unloading is readily available from the experiment data
by the equation:
et Dt  et
e_ t
Dt
where Dt is a small change in time and the e_ c t of the pillar
at the onset of unloading was found to have a positive value of 1.9  103 s1, which is much larger than the value of e_ e t and e_ th t. Therefore, signicant creep occurs in
the deformation of duralumin micropillars.

R. Gu, A. H. W. Ngan / Scripta Materialia 68 (2013) 861864

The microstructures in the 10 1% deformed duralumin micropillars aged at RT and under peak conditions
with dierent sizes are shown in Figure 4(ae). Here, the
TEM images were taken with diraction vectors g under
which visible dislocations were most abundant. Figure 4(a) shows that the dislocations in an RT-aged 1 lm
pillar distributed inhomogeneously within the specimen.
In some regions they were entangled together with density
up to 1  1015 m2 as measured by a line-intercept
method [14], as shown in Figure 4(b), while in other regions, such as a slip band, the dislocation density was
much lower, indicating that a lot of the dislocations had
escaped from the pillar. Large dispersoids are also visible
as dark round spots with sizes in the submicron range. In
the typical TEM image of an RT-aged 5 lm pillar shown
in Figure 4(c), the dislocation distribution is similar to
that in the 1 lm pillar shown in Figure 4(b). However,
for the peak-aged pillars in Figure 4(d and e), the size effect on the dislocation distribution is more signicant
the dislocation density in the 1 lm pillar in Figure 4(d)
is 1  1014 m2, which is signicantly smaller than the
2  1015 m2 in the 5 lm pillar in Figure 4(e).
Recent research [2427] has revealed that, in an Al
CuMg alloy, the rst stage of hardening after the water
quench is due to the formation of CuMg co-clusters at
low temperature, and after articial aging a hardness
plateau is reached when the S phase and GuinierPrestonBagaryatsky (GPB) zones co-exist, with the S phase
(Al2CuMg) being the major strengthening precipitate,
while the rod-shaped dispersoids of T phase (Al20Cu2Mn3) commonly found in duralumin do not make a
signicant contribution towards hardening. In the
peak-aged state the lath-shaped S precipitates have submicron size and spacing [24,25], which are much larger
than the nanometer-sized co-clusters and GPB zones
produced during RT aging [25].
Figure 2 shows that the 2% proof strength of duralumin
micropillars decreases with size in te power law r / Dm ,
with m 0.34 and 0.51 for the RT-aged and peak-aged
states, respectively. The value of m for duralumin pillars
here is therefore similar to that in pure Al pillars with sig-

Figure 4. (a) TEM montage of a longitudinal section of an 1 lm RTaged pillar. (b, c) High-magnication bright-eld images of RT-aged
pillars of 1 and 5 lm diameter, respectively. (d, e) High-magnication bright-eld images of peak-aged pillars of 1 and 5 lm
diameter, respectively. All pillars were deformed to 10 1%.

863

nicant pre-straining, which decreases from 0.98 with no


pre-straining to 0.51 with 15% pre-straining [12]. The results here indicate that, in situations where the internal
microstructural length scale impeding dislocation motion
is very ne, whether it is the spacing between the precipitates, as in duralumin, or that between initial dislocations,
as in highly pre-strained Al, the size eect of strength will
become much milder. In such cases, the internal microstructural length scale, rather than eects controlled by
the extrinsic specimen size, dominates the resistance to dislocation motion and hence strength. As mentioned above,
the present peak-aged duralumin contains S precipitates
with submicron spacing [24,25], and in pure Al after 15%
pre-straining the spacing of the resultant dislocations is
also of a similar magnitude [12], so it is not surprising that
peak-aged duralumin and pre-strained Al pillars exhibit
very similar m values of 0.51 in the proof stress regime.
As for the RT-aged micropillars, the internal length scale
is determined by even smaller precipitates of GPB zones
and CuMg co-clusters [25], so, compared with the
peak-aged pillars, the internal length scale controlling dislocation resistance is much smaller in the RT-aged pillars,
and the external size eect on the proof strength of pillars
is even weaker, with m  0.34.
Figure 1 shows that the deformation of duralumin
pillars is more homogeneous than pristine Al pillars of
the same size [7,14]. Again, the retaining of dislocations
in the duralumin pillars by the precipitates is thought to
be the reason. The accumulation of dislocations in
duralumin pillars increases the mutual dislocation interactions, which then stabilize the deformation. Previous
strategies to retain dislocations, such as surface coating
or introducing a grain boundary, were also found to
lead to smoother and more stable deformation compared to the pristine state [1315,28,29].
Figure 4 shows that, at 10 1% strain, the dislocation density in a 5 lm peak-aged pillar (2  1015 m2,
Fig. 4(e)) is more than an order of magnitude higher
than that in a 1 lm peak-aged pillar (1  1014 m2,
Fig. 4(d)), while for the RT-aged pillars, the residual dislocation density does not vary signicantly with the pillar size and the value (1  1015 m2, Fig. 4(b and c)) is
between the former two for the peak-aged pillars. These
dierences in dislocation densities can also be understood from the precipitate types in the two aged states.
In the peak-aged 1 lm specimen, since the S precipitates
present have a rather large submicron spacing compared
to the pillar size, the mobile dislocations have a greater
chance of moving to the free surface without encountering pinning precipitates, so the residual dislocation density is lower, at 1  1014 m2. For a peak-aged 5 lm
pillar, the mobile dislocations need to travel longer distances in order to reach the free surface, so they are
more likely to be pinned up by S precipitates and the
residual density is much higher, at 2  1015 m2. This
residual dislocation density is even higher than that of
the RT-aged pillar of the same size (1  1015 m2),
as the S precipitates in the peak-aged state are much
stronger than the nanometer-sized co-clusters and
GPB zones in the RT-aged state, and hence can hold
more dislocations within the pillar volume. It is for this
reason that the bulk can exhibit the maximum strength
in the peak-aged condition.

864

R. Gu, A. H. W. Ngan / Scripta Materialia 68 (2013) 861864

The dislocation densities at 10 1% deformation observed from Figure 4 also correspond to the 10% ow
stresses in Figure 2 rather well, in that the 1 lm peak-aged
pillars exhibit the greatest strength while the 5 lm peakaged pillars exhibit the least strength, and the strength of
the RT-aged pillars is in-between, with almost no size
dependence. The 1 lm peak-aged pillars are strongest because of the partial loss of dislocations to the free surface,
i.e. according to the Orowan equation e_ qbv, the mobile
dislocation density is continuously lower so that a higher
stress is needed to make the limited quantity of mobile dislocations travel faster so as to maintain a given strain rate.
On the other hand, the larger 5 lm peak-aged pillars are
the weakest because more dislocations are now retained
inside the pillar by the precipitates, and the RT-aged pillars are of intermediate strength because of the intermediate capability of the nanoscale precipitates to retain
dislocations. As mentioned above, this inverse trend of
dislocation density vs. strength tallies well with the Orowan equation, but is notpconsistent
with Taylors work
hardening rule of r / q, suggesting that strength in
duralumin pillars is not controlled by mutual dislocation
interactions but by dislocation drag by precipitates. This
phenomenon is analogous to a previous observation
[12,14] that pre-straining can produce softening in small
(1 lm) Al pillars the pre-straining in that case increases
the mobile dislocation density, so there are now more carriers for slip.
It is also interesting to see from Figure 2 that the 2%
proof strength of the 5 lm pillars in the peak-aged condition is substantially lower than that in the RT-aged condition, but this dierence narrows down tremendously for
the 10% ow stress. In fact, even in the peak-aged state,
the bulk hardness (125 Hv) is only marginally higher
than that in the RT-aged state (110 Hv) and, since the
Vickers hardness scale is a measure of the ow stress up
to 7% strain [30], the bulk hardness should better be
compared with the 10% ow stress data than the 2% proof
stress data in Figure 2. Assuming the Vickers hardness is
3.3 times the ow strength, as is typical for hardening
materials, the bulk ow strength should therefore be 387
and 333 MPa for the peak-aged and RT-aged states of
the present duralumin, respectively, and these values are
comparable with the 10% ow stress of the 5 lm pillars
in Figure 2. The strength of the bulk state should be controlled by long-range undulations in the distribution of
the precipitates which are not present in the micron-sized
pillars, though further work is necessary to clarify the relationship between the strengths of micron-sized samples
and bulk samples.
Another remarkable observation is that signicant
creep occurs in duralumin micropillars but not in pure
aluminum pillars [7,12,14]. This is likely to be due to
the various obstacles for dislocation motion in duralumin which are not present in pure aluminum, including
interactions with precipitates and solute drag. These
obstacles are bypassed by thermal uctuations and/or
climb of the dislocations and, as a result, the resultant
deformation has a signicant viscous character. Again,
the creep behavior of microsized alloys warrants further
investigation.
To conclude, precipitate-hardened duralumin alloy
pillars exhibit a power law (r / Dm ) size dependence

of strength, as do pure Al micropillars, though the size


eect is much milder than in pure Al. In the ow stress
regime, m can drop to 0.07, which indicates no eective size dependence. The results suggest that a ne
microstructural length scale due to the precipitates suppresses the size dependence of the strength. TEM characterization reveals that the strength in these alloy
pillars is controlled by the availability of mobile dislocations vs. the viscous drag by the precipitates. Creep is
also found to be signicant, which again is indicative
of the viscous drag eects by the precipitates.
We thank the Electron Microscope Unit of
HKU for their assistance. The work described in this paper was supported by funding from the Research Grants
Council (Project No. 7159/10E) and the University
Grants Committee (Project No. SEG-HKU06) of the
Hong Kong Special Administrative Region, as well as
from the Kingboard Professorship Endowment.
[1] D.M. Dimiduk, M.D. Uchic, T.A. Parthasarathy, Acta
Mater. 53 (2005) 4065.
[2] R. Dou, B. Derby, Scripta Mater. 61 (2009) 524.
[3] J.R. Greer, J.T.M. De Hosson, Prog. Mater. Sci. 56
(2011) 654.
[4] M.D. Uchic, D.M. Dimiduk, J.N. Florando, W.D. Nix,
Science 305 (2004) 986.
[5] D.M. Dimiduk, C. Woodward, R. LeSar, M.D. Uchic,
Science 312 (2006) 1188.
[6] J.R. Greer, W.D. Nix, Phys. Rev. B 73 (2006).
[7] K.S. Ng, A.H.W. Ngan, Acta Mater. 56 (2008) 1712.
[8] Z.W. Shan, R.K. Mishra, S.A.S. Asif, O.L. Warren,
A.M. Minor, Nat. Mater. 7 (2008) 115.
[9] R. Maass, M.D. Uchic, Acta Mater. 60 (2012) 1027.
[10] T.A. Parthasarathy, S.I. Rao, D.M. Dimiduk, M.D.
Uchic, D.R. Trinkle, Scripta Mater. 56 (2007) 313.
[11] D.M. Noreet, D.M. Dimiduk, S.J. Polasik, M.D. Uchic,
M.J. Mills, Acta Mater. 56 (2008) 2988.
[12] R. Gu, A.H.W. Ngan, J. Mech. Phys. Solids (2012),
http://dx.doi.org/10.1016/j.jmps.2012.10.002.
[13] K.S. Ng, A.H.W. Ngan, Acta Mater. 57 (2009) 4902.
[14] R. Gu, A.H.W. Ngan, Acta Mater. 60 (2012) 6102.
[15] K.S. Ng, A.H.W. Ngan, Philos. Mag. 89 (2009) 3013.
[16] A. Kunz, S. Pathak, J.R. Greer, Acta Mater. 59 (2011)
4416.
[17] X.X. Chen, A.H.W. Ngan, Scripta Mater. 64 (2011) 717.
[18] A. Rinaldi, P. Peralta, C. Friesen, K. Sieradzki, Acta
Mater. 56 (2008) 511.
[19] X.W. Gu, C.N. Loynachan, Z. Wu, Y.-W. Zhang, D.J.
Srolovitz, J.R. Greer, Nano Lett. 12 (2012) 6385.
[20] D. Jang, C. Cai, J.R. Greer, Nano Lett. 11 (2011) 1743.
[21] K.A. Afanasyev, F. Sansoz, Nano Lett. 7 (2007) 2056.
[22] D. Jang, X. Li, H. Gao, J.R. Greer, Nat. Nanotechnol. 7
(2012) 594.
[23] G. Feng, A.H.W. Ngan, J. Mater. Res. 17 (2002) 660.
[24] T.S. Parel, S.C. Wang, M.J. Starink, Mater. Des. 31
(2010) S2.
[25] S.C. Wang, M.J. Starink, Int. Mater. Rev. 50 (2005) 193.
[26] S.C. Wang, M.J. Starink, Acta Mater. 55 (2007) 933.
[27] M.J. Starink, S.C. Wang, Acta Mater. 57 (2009) 2376
2389.
[28] A.T. Jennings, C. Gross, F. Greer, Z.H. Aitken, S.W. Lee,
C.R. Weinberger, J.R. Greer, Acta Mater. 60 (2012) 3444.
[29] S.-W. Lee, A.T. Jennings, J.R. Greer, Acta Mater. (2013),
http://dx.doi.org/10.1016/j.actamat.2012.12.008.
[30] K.L. Johnson, Contact Mechanics, Cambridge University
Press, Cambridge, 1985.

Vous aimerez peut-être aussi