Anda di halaman 1dari 10

Available online at www.sciencedirect.

com

Acta Materialia 60 (2012) 61026111


www.elsevier.com/locate/actamat

Eects of pre-straining and coating on plastic deformation of aluminum


micropillars
R. Gu, A.H.W. Ngan
Department of Mechanical Engineering, The University of Hong Kong, Pokfulam Road, Hong Kong, Peoples Republic of China
Received 5 June 2012; received in revised form 18 July 2012; accepted 19 July 2012
Available online 23 August 2012

Abstract
The plastic deformation of micropillars is known to be aected by whether dislocations can escape easily from the material volume,
and the extent to which the dislocations mutually interact during the deformation. In this work, pre-straining and coating are used to
modify the initial dislocation content and the constraints on the escape of dislocations. Aluminum micropillars in the size range from 1
to 6 lm, with or without thin coating by tungsten deposition and pre-straining by 7%, were compressed using a at-punch nanoindenter to study their plasticity behavior. The results reveal very dierent behavior between the size regime of a few microns and that of
1 lm. For pillars a few microns large, coating leads to signicant strengthening, and pre-straining by 7% also produces a mild strengthening eect. The proof strength also exhibits good correlation with the square root of the residual dislocation density measured by transmission electron microscopy after deformation, indicating that strength in this size regime is controlled by dislocation interactions as in
traditional Taylor hardening. Coating evidently helps retain dislocations inside the pillar, and pre-straining increases the initial dislocation content; both eects lead to more severe strain hardening during deformation. For smaller pillars 1 lm in size, however, pre-straining results in softening, although coating still leads to strengthening, and the strength exhibits no correlation with the residual dislocation
density, which remains close to the initial value even with coating. These suggest dislocation starvation in this small size regime and that
strength is controlled by the availability of mobile dislocations. Coating cannot eectively trap dislocations inside the pillar, but can still
strengthen the pillar, presumably because dislocation nucleation is more dicult at the coated surface.
2012 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Keywords: Plastic deformation; Compression tests; Dislocations; Nanoindentation

1. Introduction
The last decade or so has seen tremendous interest in the
yielding of micropillar forms of metals. In addition to their
power-law, size-dependent strength [111], these micropillars also deform in a jerky manner with continuous occurrence of strain bursts in a stochastic manner [1,2,1217]. To
explain these unusual behaviors, several theories have been
proposed. One important concept is dislocation starvation [13,15], which refers to the easy zipping of dislocations through the small crystal without accumulation and
multiplication, therefore resulting in the crystal staying in
Corresponding author.

E-mail address: hwngan@hku.hk (A.H.W. Ngan).

a continuously dislocation-starved state. Such a mode of


deformation has been proven by direct transmission electron microscopy (TEM) observations in nanometric metal
volumes [15,17]. Another group of proposed models
includes source truncation [18,19] and exhaustion hardening [20,21], which describe the jerky and stochastic nature of deformation as the result of infrequent interaction
events between the small content of dislocations present
in the microcrystal, so that the mean-eld conditions as
in the bulk state cannot be met.
Ng and Ngan [22] have demonstrated that by coating the
free surface or lling an internal cavity of aluminum micropillars with tungsten deposition, dislocations are trapped
inside the aluminum and their greatly enhanced multiplication and interactions result in smooth deformation with

1359-6454/$36.00 2012 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.actamat.2012.07.048

R. Gu, A.H.W. Ngan / Acta Materialia 60 (2012) 61026111

ultra-high proof strength and strain-hardening rate. These


results are consistent with an earlier short report [23] on
compressive behavior of Al2O3-coated gold micropillars.
Recently, Jennings et al. [24] also observed that copper
nanopillars with TiO2/Al2O3 coatings not only exhibited
much higher strength and the Bauschinger eect as in metallic thin lms with a passivation layer [2527], but also significant stochastic nature in the compression behavior. These
eects of trapping dislocations have also been veried by
dislocation dynamics simulations [28,29]. On the other
hand, attempts have also been made to understand the
eects of altering the initial dislocation contents in micropillars by pre-straining the metal prior to pillar machining,
which was found to produce a softening eect in general
[3032]. The evidence so far has therefore pointed to very
remarkable eects on the plastic deformation of micropillars
if their dislocation contents are conned as in the coating
experiments, or increased a priori as in the pre-straining
experiments. In this work, we investigate the coupled eects
of both means for conditioning the dislocation contents in
aluminum micropillars over a range of sizes in the micron
regime.
2. Experimental details
A piece of ultrapure aluminum (purity >99.99999 wt.%)
was cut from a rod of 12 mm diameter. This sample was
annealed at 500 C for 24 h, and then treated by mechanical polishing and electropolishing in a 1:9 mixture of perchloric acid and methanol at 30 C and 16 V for
3.5 min. A large grain with orientation [516] and diameter
>2 mm was detected by electron back-scattered diraction
(EBSD) and was selected for the subsequent experiments
reported below. The bulk yield strength of the grain was
estimated as a third of its Vickers hardness as assessed by
microindentation just on that grain.
Micropillars were fabricated by focused ion-beam (FIB)
milling in a Quanta 200 3D dual beam FIB/SEM system
operating at 30 kV ion beam voltage, through a series of
concentric annular pattern millings with the current varied
from 20 nA for initial coarse milling to 50 pA for nal ne
milling. In this study, three micropillar diameters of 5.6, 3.3
and 1.2 lm, all with height-to-diameter ratios of around
4:1, were fabricated. Their appearances before and after
compression were imaged by scanning electron microscopy
(SEM) in a LEO1530 microscope. A batch of the fabricated pillars was then coated by tungsten deposition in
the same FIB system on the side surface. In this procedure,
hexacarbonyl tungsten gas W(CO)6 [33] is injected near the
specimen, and this is then decomposed by the gallium ion
beam into a tungsten-rich solid deposit on the specimen.
This deposit is not pure metallic tungsten but is alloyed
with a signicant amount of Ga and is in a nanocrystalline
clustered state [33]. For the aluminum specimens in this
work, the ion current used for tungsten coating was
<0.5 nA, and the coating thickness was adjusted by a nal
milling step to achieve volume fractions of 520%.

6103

Compression of the micropillars was performed in a


G200 nanoindenter with a at-ended diamond punch at
room temperature. The at punch was manufactured from
a diamond Berkovich tip by FIB milling, and it has diameter and height of around 8.5 and 1.5 lm, respectively. The
compression tests were performed in a load-controlled
manner, at a loading rate of 40 lN s1 for pillars larger
than 3 lm in diameter, or 10 lN s1 for smaller pillars, in
order to maintain the strain rate of all the samples at
103 s1 as the deformation behavior for both the
uncoated and coated micropillars can be strain-rate dependent [24,34,35].
Representative deformed pillars were thinned down by
FIB milling to electron transparency for TEM examination. The TEM samples were longitudinal sections of the
deformed pillars made by milling away two sides of the pillars along their longitudinal direction, and the orientation
of dierent samples in the selected Al grain were maintained the same, so that comparable diraction conditions
could be achieved in the TEM. The specimens were then
welded onto an in situ Omniprobe by tungsten deposition
and cut o from the bulk base by ion milling. The slices
were transferred to standard TEM copper grids by a similar welding and cutting process. Finally the specimens were
thinned further to 150 nm thickness for electron transparency. TEM examination was carried out in a Philips Tecnai
microscope operating at 200 kV.
After compression and characterization of the pillars
fabricated on the annealed bulk sample, the latter was cold
rolled to 7% reduction. Fresh micropillars and tungstencoated pillars with similar sizes were fabricated on the same
grain as before, and these specimens, now with a 7% prestrain, were subjected to compression testing and TEM
analysis using the same procedures as above. This procedure enabled pre-strained micropillars of the same crystallographic orientation as the pristine pillars to be fabricated
and investigated.
3. Results and analysis
3.1. Deformation behavior of tungsten-coated specimens with
pre-straining
Fig. 1 shows the nominal stressnominal strain curves of
compression tests on the 7% pre-strained pillars of dierent
sizes with and without tungsten coating. For the coated pillars, the volume fraction Vw of the tungsten coating was
varied from 7% to 20% in the three sets of experiments
shown. The typical appearances of the deformed pillars
are shown on the right of each group. The stressstrain
responses of both coated and uncoated micropillars without pre-straining have been studied [22,24] and, as mentioned above, coating can signicantly raise the strainhardening rate and smoothen the stressstrain response.
For the pre-strained cases as shown in Fig. 1, compared
with the uncoated specimens of similar sizes, strengthening
and smoothening of the stressstrain curves can also be

6104

R. Gu, A.H.W. Ngan / Acta Materialia 60 (2012) 61026111

Fig. 1. Normal stressnormal strain curves of 7% pre-strained aluminum pillars of diameter (a) 5.6 lm, (c) 3.3 lm and (e) 1.2 lm. Colored curves (online
version only) are for coated pillars with dierent volume fractions Vw of the tungsten coating; black curves are for uncoated pillars of similar sizes.
Representative SEM images of the deformed coated pillars with diameter of (b) 5.6 lm, (d) 3.3 lm and (f) 1.2 lm are also shown. (For interpretation of the
references to color in this gure legend, the reader is referred to the web version of this article.)

achieved with tungsten coating of the pillars, and the eects


are not sensitive to the volume fraction of the tungsten
coating. However, for all the three specimen sizes, sudden

softening manifested as a plateau in the stressstrain curve


always occurs at a later stage of deformation, and similar
events also happen without pre-straining [22]. For the

R. Gu, A.H.W. Ngan / Acta Materialia 60 (2012) 61026111

5.6 and 3.3 lm coated pillars, Fig. 1a,c indicate that the
deformation is smooth with continuous strain-hardening
up to 5% strain, but at larger strains the strain-hardening
ceases and a plateau follows. For the smaller 1.2 lm pillars in Fig. 1e, the softening at 5% strain is so rapid that it
corresponds to a huge and uncontrollable strain burst of
10%. The post-deformation SEM images of the coated
pillars shown in Fig. 1b, d and f reveal the occurrence of
only one or two intensive slip bands, around which the
tungsten coating shows signs of cracking and aking o
the pillar surface. As for the 1.2 lm pillars, which deformed
with a large avalanche, the post-deformation SEM morphology in Fig. 1f shows an intensive shear step corresponding to almost complete shear fracture of the pillar
into two segments. The large strain burst is thought to be
related to the cracking of the coating at a large stress
[22,24], and as this happens substantial slip of the pillar
core would proceed at the cracked region.
3.2. Strengthening of tungsten-coated specimens
The stress plotted in Fig. 1a, c and e is the nominal stress
r, which is the applied load divided by the gross cross-sectional area of the coated pillars including the coating thickness. To obtain the stress rAl applied onto the aluminum
core of the coated pillars, a composite model is used as
in Ref. [22]:
r rAl V Al rw V w ;
where VAl and Vw are the aluminum and tungsten deposition volume fractions, respectively, and rw is the stress sustained by the tungsten coating. The intrinsic strength of the
tungsten coating deposition fabricated by the present FIB
system was investigated previously [22], and its 2% proof
strength was found to be 100 MPa. Hence the 2% proof
strength rAl of the aluminum core can be calculated from
the overall nominal stress r using the above rule of mixtures. Such a rule-of-mixtures calculation assumes simple
load sharing between the constituent phases without mutual interactions, and the objective here is indeed to see
whether the pillar cores intrinsic strength can be explained
by such an eect, since if this is the case, the calculated rAl
would agree with the strengths of the uncoated pillars. The
results for dierent cases, including the uncoated pillars
with no pre-straining (i.e. the pristine group), are shown
in Fig. 2. For the coated pillars in both pre-straining conditions, their rAl data calculated from the rule of mixtures
are always higher than the strengths of the uncoated pillars
of the same size in accordance with previous ndings [22],
implying that the strength of the pillar core is aected by
some interaction eects from the coating. However, the
coated and pre-strained pillars (the red dotted line in
Fig. 2) are substantially weaker than the coated pillars with
no pre-straining (the black dotted line) at sizes of 1.2 lm:
the 2% proof strength of the coated/pre-strained group at
1.2 lm size is 172.6 24.1 MPa which is only about
54% of the coated group without pre-straining. At sizes

6105

Fig. 2. Average 2% proof strength of aluminum micropillars with and


without coating and pre-straining.

of 3.3 lm, the strength of the coated/pre-strained group


is 133.2 8.8 MPa which is 78% of that of the coated
group without pre-straining, and at 5.6 lm, the strength
of the coated/pre-strained group is 101.4 16.1 MPa
which is 1.1 times that of the coated pillars without prestraining. Pre-straining and coating evidently result in different eects on the strength data in Fig. 2 between large
(>3.3 lm) and small (1 lm) pillar sizes. For the uncoated
specimens (the two solid curves in Fig. 2), their strength
varies with diameter almost following a power law, and
the pre-strained specimens exhibit a smaller power exponent than the pristine case; this phenomenon will be examined in greater detail, separately [36].
3.3. Dislocation distribution in the specimens before
compression
TEM characterization was carried out on representative
deformed specimens of each group in order to understand
the dislocation distributions in the micropillars. First it is
necessary to study the dislocation distribution in the initial
states before pillar fabrication. Fig. 3a and b are TEM
montages of the annealed and 7% pre-strained states of
the aluminum bulk from which the micropillars were fabricated. These TEM specimens were prepared by wire-cutting thin slices from the bulk piece and electropolishing
using the traditional twin-jet method. It can be seen in
Fig. 3a that in the annealed bulk dislocation lines distribute
rather separately without extensive clustering, whereas in
the 7% pre-strained state as in Fig. 3b the dislocation lines
are much denser and assemble into subcells. The dislocation density in the two states was measured by a line-intercept method, i.e. the average distance d between
dislocations was measured from the number of intersection
points random lines drawn on the TEM image would make
with the dislocations, and the dislocation density q was
estimated as 1/d2. The dislocation density q for the
annealed state is estimated as 9.0  1011 m2, and that
for the 7% pre-strained state is 2.2 times higher at
2.0  1012 m2. Vickers hardness (HV) measurement of

6106

R. Gu, A.H.W. Ngan / Acta Materialia 60 (2012) 61026111

Fig. 3. (a and b) TEM montages of annealed and 7% cold reduction aluminum samples [36] prepared by the twin-jet method. (c) TEM image of a FIBprepared specimen from the annealed bulk taken near the [1 0 1] pole.

Fig. 4. (a and c) Montages of TEM images of the longitudinal sections of 5.6 lm uncoated pillars without (a) and with (c) 7% pre-strain. The top of the
pillar is towards the right side (the same orientation is used for all subsequent TEM images). (b and d) Higher-magnication TEM images of each
condition.

the annealed and pre-strained states was performed and


this gives yield stress estimates of the two states as 4.5
and 5.8 MPa, respectively. From the Taylor hardening
p
equation r / q, the dislocation density in the prestrained state should be about 1.7 times that of the
annealed state, and this ratio is close to the 2.2 from direct
measurement of the dislocation densities.
The assessment of the dislocation content in the
deformed pillars, to be reported below, was carried out
on FIB-thinned longitudinal sections of the pillars as
described in Section 2. Therefore, irrespective of the dislocation structures as seen from twin-jet electropolished
TEM samples, the eects of ion bombardment during
the FIB process need to be assessed. Fig. 3c shows a
TEM image of a FIB-produced section of an undeformed

pillar without pre-straining, and the dislocation density in


this specimen is estimated to be 1.0  1013 m2, i.e. an
order of magnitude higher than that seen in the twin-jet
prepared sample (cf. Fig. 3a). It should be noted that
the FIB-fabricated micropillars were much thicker than
the TEM foils, and so the background dislocation density
of 1013 m2 as seen from Fig. 3c may not be representative of the interior of the micropillars. Furthermore, any
dislocation density orders of magnitude higher than
1013 m2 found in a compressed pillar should be interpreted as arising from compression process itself, rather
than due to the FIB damage during the TEM sample
preparation. These two points should be borne in mind
when interpreting the dislocation density measurements
reported below.

R. Gu, A.H.W. Ngan / Acta Materialia 60 (2012) 61026111

6107

Fig. 5. (a and c) Montages of TEM images of the longitudinal sections of 1.2 lm uncoated pillars without (a) and with (c) 7% pre-strain. (b and d)
Higher-magnication TEM images of each condition.

Fig. 6. (a and c) Montages of TEM images of the longitudinal sections of 5.6 lm tungsten coated pillars without (a) and with (c) 7% pre-strain. (b and d)
Higher-magnication TEM images of each condition.

3.4. Dislocation distribution in the deformed specimens


Figs. 47 show TEM microstructures of representative
pillars in each condition after compression, and details are
given Table 1. The dislocation densities are estimated as

the average values from several high-magnication TEM


images taken with diraction vectors g under which visible
dislocations were most abundant. Although the deformation strains of the specimens are dierent, experimental
observations have showed that strain is not an important

6108

R. Gu, A.H.W. Ngan / Acta Materialia 60 (2012) 61026111

Fig. 7. (a and c) Montages of TEM images of the longitudinal sections of 1.2 lm tungsten coated pillars without (a) and with (c) 7% pre-strain. (b and d)
Higher-magnication TEM images of each condition.

Table 1
Details of deformed aluminum micropillars for TEM characterization.
Diameter
(lm)

Pre-strain

Tungsten volume
fraction

Max. stress
(MPa)

Final strain

5.6
5.6
1.2
1.2
5.6
5.7
1.1
1.2

None
7%
None
7%
None
7%
None
7%

None
None
None
None
12%
20%
10%
20%

58
65
240
209
172
227
333
295

0.05
0.11
0.07
0.10
0.06
0.19
0.04
0.04

factor inuencing the dislocation densities in deformed


microcrystals [20,37]. Fig. 4 compares the dislocation distribution with and without the 7% pre-straining in uncoated
pillars of 5.6 lm diameter. No obvious subcell structure
could be seen in the case without pre-straining (Fig. 4a),
and the dislocation density is estimated to be 1.0  1014 m2,
i.e. an order of magnitude higher than the 1013 m2
background introduced by the FIB process itself. As for
the pre-strained case, dislocation subcells several microns
large are discernible in the montage shown in Fig. 4c, and
it may be due to the fact that this pillar was fabricated from
the cold-rolled bulk which already contained dislocation
cellular structures as shown in Fig. 3b. From the highermagnication image shown in Fig. 4d, it is clear that dislocations are more abundant than the case with no pre-strain
(Fig. 4b), and their density is about 3.0  1014 m2.

g for dislocation density


estimation
1 1 1; 0 2 0
2 0 0; 1 1 1
1 1 1
2 0 0; 1 1 1

1 1 1; 1 1 1
1 1 1; 1 1 1
1 1 1
1 1 1

Corresponding
gures
Fig.
Fig.
Fig.
Fig.
Fig.
Fig.
Fig.
Fig.

4a and b
4c and d
5a and b
5c and d
6a and b
6c and d
7a and b
7c and d

Fig. 5 compares the dislocation arrangements in the


smaller size of 1.2 lm pillars. For both cases with and
without pre-straining, the dislocations arrange randomly
with density 2.8  1013 m2 for the case without pre-strain,
and 5.4  1013 m2 with pre-strain, i.e. both values are signicantly smaller than those of the larger pillars under similar conditions, and are close to the 1013 m2 background in
the initial states. Pre-straining evidently results in no obvious eect on the dislocation distribution in the small pillars.
Figs. 6 and 7 show the dislocation structures in the
coated micropillars of large and small sizes, respectively.
Fig. 6 shows that after deformation, dislocation cellular
structures were developed in the coated 5.6 lm pillars
both with or without pre-strain. The higher-magnication
TEM images in Fig. 6b and d show that dislocations in
the subgrains were abundant at density 7.0  1014 m2

R. Gu, A.H.W. Ngan / Acta Materialia 60 (2012) 61026111

Fig. 8. Estimated dislocation densities in the deformed micropillars.

for the case without pre-strain, or 1.0  1015 m2 with prestrain. On the contrary, for the 1.2 lm coated pillars, no
dislocation cell could be observed in the TEM montages
shown in Fig. 7a and c with or without pre-strain. In these
samples, dislocations are scarce and arrange in a random
manner, and their densities are estimated as 5.0  1013 m2
without pre-strain and 7.0  1013 m2 with pre-strain, i.e.
still very close to the background value of 1013 m2.
Fig. 8 summarizes the measured dislocation densities of different types of pillars after compression. It shows that in the
case of the 5.6 lm pillars, coating and pre-straining can
result in a signicant increase in the dislocation density during deformation. On the contrary, in the small 1.2 lm pillars, the dislocation content does not vary much with or
without coating or pre-straining, and stays at the low magnitude of 1013 m2, comparable to the initial state prior to
compression.
4. Discussion
The present results in Fig. 2 indicate that coating the
aluminum micropillars always leads to strengthening irrespective of the sizes of the pillars within the range studied,
and whether or not pre-straining was applied. However,
the eect of the 7% pre-straining is very dierent between
large and small pillars. For larger 5.6 lm pillars, prestraining results in a small strengthening eect in both
the coated or uncoated conditions, but for the small
1 lm pillars, pre-straining leads to softening in both the
coated and uncoated conditions, and the eect on the
coated condition is larger. The results in Fig. 8 also show
that at the small size of 1.2 lm, the dislocation content
remains close to the initial state in all groups, but this is
not so in the large pillar size of 5.6 lm. These indicate
that the dislocation mechanisms are very dierent between
the large and small pillars.

6109

Fig. 9. Strength vs. square root of residual dislocation density for the
5.6 lm pillars.

size, the strength data in Fig. 2 are in the following descending order: coated with pre-strain > coated without pre-strain > uncoated with pre-strain > uncoated without prestrain, and the dislocation density data in Fig. 8 follow
exactly the same order. Hence, for the 5.6 lm pillars,
strength evidently correlates with the residual dislocation
density. Fig. 9 plots the strength data vs. the square root
of the residual dislocation density data for the 5.6 lm pillars, and a proportionality relation between these two quantities is apparent. Hence, the natural conclusion for this size
regime is that strength is controlled by mutual dislocation
interactions as in traditional Taylor hardening, i.e.
p
r / q. This conclusion veries the key assumption of Taylor hardening in a number of theories on size eect of
strength [18,19,36], as well as the prediction of dislocation
dynamics simulations [21]. In this larger size regime, coating
produces strengthening because it traps dislocations inside
the pillar and their frequent interactions lead to multiplication, so that the density of mobile dislocations is maintained
at a high level [22]. Without coating, some dislocations can
annihilate at the pillars free surface, but since the pillar size
is not small, not all dislocations can travel freely to the free
surface without interaction with other dislocations. The
result is that, even without coating, the residual dislocation
density is still higher than the initial state with signicant
production. Pre-straining by 7% in this size regime evidently
results in a mild increase in the initial dislocation content
(cf. Fig. 3a and b), and indeed in Fig. 8, the residual dislocation density is higher with pre-straining in both the coated
and uncoated case. The higher initial dislocation content
from pre-straining leads to a small strengthening eect via
the Taylor mechanism as shown in Fig. 2, both for the
coated and uncoated case. In all, in this larger size regime,
the eects of coating and pre-straining on strength and
residual dislocation density can be understood as arising
from the Taylor hardening mechanism.

4.1. The larger size (>3.3 lm) regime


4.2. The small size (1 lm) regime
Comparison between Figs. 2 and 8 in fact shows that at
the 5.6 lm pillar size, the strength follows the same trend
as the residual dislocation density after deformation. At this

The most remarkable phenomenon to note at the small


pillar size of 1.2 lm is that the residual dislocation

6110

R. Gu, A.H.W. Ngan / Acta Materialia 60 (2012) 61026111

density in Fig. 8 has not grown much from the initial low
value, but the strength data in Fig. 2 vary by a large extent
depending on whether the pillar is coated or pre-strained.
There is therefore no correlation between strength and
the residual dislocation density, i.e. the results here suggest
that strength in this regime is not controlled by the Taylor
mechanism. Another important phenomenon from Fig. 2 is
that pre-straining softens the pillars in both the coated and
uncoated groups at this small size, and this is in fact opposite to the Taylor eect where more existing dislocations
should produce strengthening. Therefore, the natural conclusion to draw for this size regime is that strength is controlled by the availability of mobile dislocations, in the
sense of the dislocation starvation concept as described
in Section 1, or simply that pertaining to the Orowan equation e_ qbm if the pillar is not completely depleted of dislocations. Pre-straining evidently results in more initial
dislocations so that strength is decreased at the same strain
rate. However, these mobile dislocations zip through the
small crystal easily, so that the residual dislocation content
stays at the same range. Strength is controlled by the mechanism(s) by which new dislocations are generated, presumably at the pillars surface.
The eects of coating are more dicult to comprehend
in this small size regime. Fig. 2 shows that coating still
strengthens the 1.2 lm pillars as in the larger pillars,
but with pre-straining the strengthening eect is less than
that without pre-straining. The persistently low residual
dislocation density even with coating means that the latter
is not eective in trapping the generated dislocations in this
small size regime. Two factors may explain this. First, the
applied stress in this size regime is now very high, and secondly, with the same volume ratio Vw of the coating it is
now very thin, compared to the larger pillars. Therefore,
the coating may fail easily when interacted upon by an
approaching dislocation, and may not be able to retain
the latter at the Alcoating interface. However, in Fig. 2,
the 1.2 lm coated pillars are always stronger than the
uncoated ones, and this is particularly so without prestraining. This seems to indicate that although the coating
cannot trap the generated dislocations, it can make the dislocation generation process itself more dicult. It is quite
possible that nucleation of dislocations at a coated surface
of the pillar is more dicult than at a free surface. The
present results therefore suggest that the strengthening
mechanism of coating is quite dierent in the small size
regime, as compared with dislocation trapping in the larger
size regime.
5. Conclusions
Eects of coating and pre-straining on deformation of
aluminum micropillars were investigated by compression
experiments using a nanoindenter with a at-ended diamond punch. The results indicate that in the larger size
regime of a few microns, the strength of the pillars correlates well with the square root of the residual density of

dislocations after compression and is therefore controlled


by Taylor-type interaction mechanisms. In this regime,
coating traps dislocations and pre-straining can further
raise the initial dislocation contents, and so both are means
of strengthening the material. In the size regime approaching 1 lm, however, the residual dislocation density remains
close to the initial value and pre-straining leads to softening, suggesting that strength is controlled by the availability of mobile dislocations. Coating cannot trap dislocations
eectively but can still produce a signicant strengthening
eect, suggesting that dislocation nucleation at a coated
surface of the pillar is more dicult.
Acknowledgements
We thank the Electron Microscope Unit of HKU for
their assistance. The work described in this paper was supported by grants from the Research Grants Council (Project No. 7159/10E) as well as from the University Grants
Committee (Project No. SEG-HKU06) of the Hong Kong
Special Administrative Region.
References
[1] Dimiduk DM, Uchic MD, Parthasarathy TA. Acta Mater
2005;53:4065.
[2] Dimiduk DM, Uchic MD, Rao SI, Woodward C, Parthasarathy TA.
Modell Simul Mater Sci Eng 2007;15:135.
[3] Volkert CA, Lilleodden ET. Philos Mag 2006;86:5567.
[4] Frick CP, Clark BG, Orso S, Schneider AS, Arzt E. Mater Sci Eng A
Struct 2008;489:319.
[5] Ng KS, Ngan AHW. Philos Mag 2008;88:677.
[6] Kim J-Y, Greer JR. Acta Mater 2009;57:5245.
[7] Lee S-W, Han SM, Nix WD. Acta Mater 2009;57:4404.
[8] Han SM, Bozorg-Grayeli T, Groyes JR, Nix WD. Scripta Mater
2010;63:1153.
[9] Kim J-Y, Jong D, Greer JR. Acta Mater 2010;58:2355.
[10] Sun Q, Guo Q, Yao X, Xiao L, Greer JR, Sun J. Scripta Mater
2011;65:473.
[11] Ye J, Mishra RK, Sachdev AK, Minor AM. Scripta Mater
2011;64:292.
[12] Uchic MD, Dimiduk DM, Florando JN, Nix WD. Science
2004;305:986.
[13] Greer JR, Oliver WC, Nix WD. Acta Mater 2005;53:1821.
[14] Dimiduk DM, Woodward C, LeSar R, Uchic MD. Science
2006;312:1188.
[15] Greer JR, Nix WD. Phys Rev B 2006;73.
[16] Ng KS, Ngan AHW. Acta Mater 2008;56:1712.
[17] Shan ZW, Mishra RK, Asif SAS, Warren OL, Minor AM. Nat Mater
2008;7:115.
[18] Parthasarathy TA, Rao SI, Dimiduk DM, Uchic MD, Trinkle DR.
Scripta Mater 2007;56:313.
[19] Rao SI, Dimiduk DM, Tang M, Parthasarathy TA, Uchic MD,
Woodward C. Philos Mag 2007;87:4777.
[20] Noreet DM, Dimiduk DM, Polasik SJ, Uchic MD, Mills MJ. Acta
Mater 2008;56:2988.
[21] Rao S, Dimiduk D, Parthasarathy T, Uchic M, Tang M, Woodward
C. Acta Mater 2008;56:3245.
[22] Ng KS, Ngan AHW. Acta Mater 2009;57:4902.
[23] Greer JR. Mater Res Soc Symp 2007 [0983-LL08-03].
[24] Jennings AT, Gross C, Greer F, Aitken ZH, Lee SW, Weinberger CR,
et al. Acta Mater 2012;60:3444.
[25] Xiang Y, Vlassak JJ. Scripta Mater 2005;53:177.

R. Gu, A.H.W. Ngan / Acta Materialia 60 (2012) 61026111


[26] Nicola L, Xiang Y, Vlassak JJ, Van der Giessen E, Needleman A. J
Mech Phys Solids 2006;54:2089.
[27] Xiang Y, Vlassak JJ. Acta Mater 2006;54:5449.
[28] Zhou C, Biner S, LeSar R. Scripta Mater 2010;63:1096.
[29] El-Awady JA, Rao SI, Woodward C, Dimiduk DM, Uchic MD. Int J
Plast 2011;27:372.
[30] Bei H, Shim S, Pharr GM, George EP. Acta Mater 2008;56:4762.

[31]
[32]
[33]
[34]
[35]
[36]
[37]

6111

Lee SW, Han SM, Nix WD. Acta Mater 2009;57:4404.


Nix WD, Lee S-W. Philos Mag 2011;91:1084.
Takeguchi M, Shimojo M, Furuya K. Jpn J Appl Phys 2007;46:6183.
Jennings AT, Li J, Greer JR. Acta Mater 2011;59:5627.
Zhu T, Li J, Samanta A, Leach A, Gall K. Phys Rev Lett 2008:100.
Gu R, Ngan AHW; submitted for publication.
Oh SH, Legros M, Kiener D, Dehm G. Nat Mater 2009;8:95.

Anda mungkin juga menyukai