Vous êtes sur la page 1sur 11

Available online at www.sciencedirect.

com

Acta Materialia 57 (2009) 16131623


www.elsevier.com/locate/actamat

Eects of specimen geometry and base material


on the mechanical behavior of focused-ion-beam-fabricated
metallic-glass micropillars
Y. Yang a,*, J.C. Ye a, J. Lu a,*, F.X. Liu b, P.K. Liaw b
a

Department of Mechanical Engineering, The Hong Kong Polytechnic University, Hung Hom, Kowloon, Hong Kong
Department of Materials Science and Engineering, The University of Tennessee, Knoxville, TN 37996-2200, USA

Received 14 August 2008; received in revised form 25 November 2008; accepted 29 November 2008
Available online 3 January 2009

Abstract
An investigation of the focused-ion-beam-based microcompression experiments was conducted using metallic-glass micropillars. The
results displayed an apparent geometry dependence of the measured pillars Youngs modulus if the formula in the literature was used for
the analysis of the experimental data. However, if the eects of the base material and pillar geometry were taken into account with the aid
of nite-element simulations, it was shown that the microcompression experiments can reach a resolution similar to that of the nanoindentation tests in the measurement of the materials mechanical properties, and therefore provide an alternative to the nanoindentation
experiment in applications that require the characterization of local mechanical properties in multi-structured/multi-phased material
systems.
2008 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Keywords: Compression tests; Nanoindentation; Metallic glasses; Finite-element analysis

1. Introduction
Microcompression experiments have recently been
widely used to study the mechanical behavior of materials
at the micrometer and submicrometer scales [110]. The
materials investigated displayed distinctly dierent
mechanical behavior, relative to their macroscopic counterparts, with the reduction of their characteristic size to the
micron and submicron scales. Thus, microcompression
studies have recently led to a renewed interest in exploring
the materials size eect on their mechanical properties.
In order to prepare specimens suitable for microcompression tests, the focused ion beam (FIB) [1,2,5,6] and
electrodeposition techniques [3,4] are usually employed.

*
Corresponding authors. Tel.: +852 2766 6652 (Y. Yang), +852 2766
6665 (J. Lu).
E-mail addresses: mmyyang@polyu.edu.hk (Y. Yang), mmmelu@
inet.polyu.edu.hk (J. Lu).

With FIB-milling, micro/nanopillars are carved out in the


surface of a substrate material; while, with electrodeposition, micro/nanopillars can be grown out of a substrate.
Once the specimens, i.e., the micro/nanopillars, are ready,
a modied nanoindentation system, tted with a at-end
diamond indenter, is used for carrying out the microcompression tests. During the engagement of the indenter, the
micropillar protruding from the substrate has the rst contact with the indenter tip and is compressed afterwards.
The loaddisplacement data are, then, continuously
recorded in the same way as in a regular nanoindentation
test [11].
In comparison with the electrodeposition-based microcompression tests, the FIB-based microcompression tests
attained greater interest and attention in the research community due to their simplicity and ease of operation. They
have been applied to materials of various kinds, such as
single crystals [14], bulk-metallic glasses [5,6], shape memory alloys [12] and other materials, which cannot be

1359-6454/$34.00 2008 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.actamat.2008.11.043

1614

Y. Yang et al. / Acta Materialia 57 (2009) 16131623

obtained easily in a form suitable for micromechanical testing. However, in spite of the successes achieved with the
microcompression tests pertaining to revealing the materials size eect, some aspects of the tests still remain elusive from an experimental viewpoint.
Due to the limitations in the fabrication process, the
shape of the FIB-fabricated micropillars is dierent from
a perfect cylinder [2,5,6,8]. Therefore, the stress and
strain in the micro/nanopillars cannot be derived with ease
as in macroscopic compression tests; neither can the pillars
Youngs modulus. These geometrical imperfections usually
comprise taper angles in the pillar diameter and rounding
at the pillar/substrate transition due to the ion-beam divergence (see the inset of Fig. 1).
As the micro/nanopillar is sitting on a substrate with
the same mechanical properties, the substrate will make
the micro/nanopillar look more compliant under compressive loading. In view of these factors, corrections were
made to interpret the experimental data in the elastic
regime for the measurements of the materials Youngs
moduli [2,5,6,15,16]. However, these corrections found in
the literature were either based on some simplifying
assumptions [2,5,6] or specic to a certain geometry and
material [15,16]. Therefore, the mechanics of how the
aforementioned geometrical imperfections and the substrate act together to aect the measurements of the
Youngs moduli of the micro/nanopillars is still not well
understood. This factor is crucial as the precise measurement of the sample Youngs modulus is usually the rst
step to justify the experimental data for further use. Moreover, given the nature of microcompression tests in probing
the materials local mechanical properties, microcompression experiments may have the potential of measuring the
mechanical properties of an individual constituent in a
multi-phased/multi-structured metallic-glass-based material, such as metallic-glass coatings, metallic-glass-based
laminates and even plasticity gradient materials, whose
mechanical properties are hard to obtain by conventional

macroscale uniaxial testing or involve the decoupling of


the structural interactions when the nanoindentation
method is used. Therefore, it is worthwhile to have a thorough investigation of the validity of the microcompression
experiments that have been used in the mechanical property measurement at the micrometer and submicrometer
scales.
In this paper, the formula that can be found in the literature is rst used to extract the Youngs modulus of the Zrbased metallic-glass micropillars fabricated via FIB. The
dimensional analysis coupled with the nite-element simulation is then employed for the systematic investigation of
the elastic deformation process of the pillar/substrate system to derive a new formula for the improved measurement
of the pillars Youngs modulus. The results are compared
to those obtained from the macroscale compression and
nanoindentation tests on the same material. The implications are discussed afterwards for the future applications
of the microcompression tests on metallic glasses and other
materials systems suitable for the microcompression study.
2. Analysis and modeling
In the literature, an approximate formula similar to
what follows was usually used for the extraction of the pillars Youngs moduli [2,5,6,15,16]. As shown in Fig. 1, setting the origin of the coordinate system at the top of the
pillar with the positive direction pointing downwards, the
total displacement, dt, of the pillar upon compression
through a at-end rigid punch can be approximated as1:
Z H
4P
dh ds
1
dt
pED0 2h tanh2
0
where P is the total compressive load applied on the pillar;
E is the pillars Youngs modulus (the same as that of the
substrate); H, D0 and h denote, respectively, the pillar
height, top diameter, and taper angle; and ds is the displacement of the substrate, which can be approximated with
Sneddons formula by assuming the pillar penetrating into
the substrate as a rigid body [17]:
ds

P 1  m
2ED0 2H tanh

where m is the Poissons ratio of the substrate (the same as


that of the pillar).
Substituting Eq. (2) into Eq. (1) and rearranging the
equation, we can obtain:


p1  mD0
4PH
3
E 1
pD0 D0 2H tanhdt
8H
with Eq. (3), the pillars Youngs modulus can be extracted
from the initial indentation loaddisplacement curve that

Fig. 1. The symbols used in describing the geometry of a micro/nanopillar


(the inset shows the prole of a real metallic glass nanopillar imaged under
SEM).

1
It is assumed that the machine compliance eect has already been
eliminated from the acquired displacement data prior to the use of Eq. (1).
For generality, Appendix A details the procedure to calibrate the machine
compliance for a microcompression test.

Y. Yang et al. / Acta Materialia 57 (2009) 16131623

1615

corresponds to the linear elastic deformation of the pillar


once the Poissons ratio is known. However, Eq. (3) is just
a crude approximation based on a few simplications: (1)
the stresses are assumed to be uniformly distributed over
each cross-section of the pillar; (2) the curvature eect at
the pillar/substrate transition is not considered and (3)
the smooth transition of stress and strain from the pillar
to the substrate is replaced by a sudden jump caused by
the rigid body assumption. In order to obtain a more accurate measurement of the pillars Youngs modulus, a more
in-depth analysis is needed regarding the elastic deformation process of the pillar/substrate system.
Based on the dimensional analysis of elastic deformation, the applied load P can be expressed as:


4P
H
q
dt
f
; h; ; m
4
2
H
D0
D0
pED0
where q is the radius of the curvature at the pillar/substrate
transition (Fig. 1) and f is a dimensionless function.
Comparing Eqs. (3) and (4), we can obtain


E
H
q
W
; h; ; m
5
E0
D0
D0
where E0 is the nominal Youngs modulus extracted based
on Eq. (3) and w is another dimensionless function, whose
values are to be evaluated.
In order to investigate the variation of E/E0 with the different parameters in Eq. (5), nite-element simulations
were carried out using the commercial package ANSYS
(ANSYS, Inc., Cannosburg, PA). A suitable substrate size
was selected through a few simulations of the substrate size
eect to ensure that the substrate boundary conditions had
no inuence on the simulated forcedisplacement curves
for the pillar/substrate system. Due to the symmetry, only
one half of the pillar/substrate geometry was analyzed. The
two-dimensional axisymmetrical elements were then used
in meshing the geometry with the highest element density
in the pillar. The mesh gradually coarsened with the
increasing distance from the pillar to optimize the computational cost and accuracy (Fig. 2).
As the rst attempt, the Poissons ratio was xed at a
constant of 0.3. Based on the nite-element simulations,
the variation of the ratio, E/E0, with dierent combinations of the aspect ratio, H/D0, the taper angle, h, and
the normalized radius of the curvature, q/D0, is shown
in Fig. 3ad. As expected, the direct implementation of
Eq. (3) led to a lower measurement of the Youngs modulus than the true value and, thus, the ratios of E/E0 were
all greater than unity. As the taper angle increased, the
areal fraction of each cross-section along the pillar height
with nonuniform stresses increased accordingly, indicative
of the increasing deviation from the assumptions that Eq.
(3) is based upon. This trend is manifested by the increasing E/E0 ratio with the increasing taper angle in Fig. 3ad.
On the other hand, as the aspect ratio, H/D0, decreased,
the substrate eect became more signicant, resulting in
TM

Fig. 2. The graded mesh used for the tapered pillar/substrate system with
a rounded pillar base in the FEM simulations.

an increasing E/E0 ratio as well. For instance, at an aspect


ratio around 12, an underestimation of 4080% of the
measured Youngs modulus could be developed from
Eq. (3). However, the increase in the radius of the curvature at the base of the micro/nanopillar relieved the stress
concentration at the pillar/substrate transition and substantially reduced the overall degree of the underestimation when the ratio of q/D0 reached above 0.5 (Fig. 3c
and d). In contrast, there was no signicant change
observed in the E/E0 ratio for 0.01 6 q/D0 6 0.1
(Fig. 3a and b).
In addition to the geometry and the associated substrate
eects, a multitude of FEM simulations were conducted to
investigate the Poissons ratio eect. The results showed
that the inuence was negligible for the Poissons ratios
ranging from 0.2 to 0.48, which spans the spectrum of
the magnitude of the Poissons ratio for most bulk-metallic
glasses [1823]. As an example, some of the results are plotted in Fig. 4. It shows that, although there is a slight
increase in E/E0 with increasing m at the aspect ratio close
to unity, it is still negligibly small, as compared to the inuences brought about by the other factors. For the micro/
nanopillars of high aspect ratios, the variation in E/E0 with
the increasing Poissons ratio is within the computational
error.
3. Experiments
In order to justify the above analyses, microcompression
studies were conducted on a series of Zr50Cu37Al10Pd3
micropillars of dierent geometries (Fig. 5), whose top radii
ranged from 0.6 to 4 lm and aspect ratios from 1.5 to
10. With the conventional elastic buckling analysis, it can
be shown that the critical aspect ratio was about 11 for the

1616

Y. Yang et al. / Acta Materialia 57 (2009) 16131623

Fig. 3. Variation of the ratio E/E0 with the combinations of H/D0 and h at dierent q/D0 for a xed Poissons ratio of 0.3.

Fig. 4. The eect of the Poissons ratio on E/E0 at dierent aspect ratios
for a xed radius of curvature and taper angle.

investigated metallic glass system (Appendix B). Therefore,


only a limited number (one out of four microcompression
tests) of useful experimental data can be obtained for the
high-aspect-ratio micropillars.
In the fabrication process of the bulk-metallic glass,
high-purity metals including Zr (99.95%), Cu (99.999%),
Al (99.999%), and Pd (99.995%), were used. The master
alloy with a nominal composition of Zr50Al10Cu37Pd3
was then obtained by arc melting. To ensure the homogeneous distribution of the chemical elements, the ingot was
melted and ipped at least ve times. The glassy alloy was
fabricated by suction-casting the master alloy into a copper mold under a Ti-gettered Ar atmosphere.
Following the procedure in Ref. [13], the micropillars
were fabricated on the surface of the bulk-metallic glass
using a dual beam scanning electron microscopy (SEM)/
FIB system (Quanta 200 3D, FEI) and the FIB fabrication
processes were performed by incrementally reducing the
ion-beam-current density to minimize the potential FIB
damage in the micropillars, which resulted in the appearance of a series of concentric rings with the base of the
micropillar at the center (Fig. 5).

Y. Yang et al. / Acta Materialia 57 (2009) 16131623

1617

Based on the SEM pictures, the taper angle and aspect


ratio can be directly measured precisely from the prole
of each micropillar. The radius of the curvature at the
micropillars base can be obtained through the geometrical relations at the pillars base (see Fig. 6), which is given
by:


p h Db  D0  2H tanh

6
q cot
4 2
2
in which Db is the bottom diameter of the micropillar, measured as the distance between the two opposite points
where the pillars meridians gradually curve into the substrate (Figs. 5 and 6). As shown in Fig. 7a, the measured
taper angles show a normal distribution in the range of
1 to 7 with an average of 3.5, which are close to
the values reported in the literature for FIB-milled micropillars [2,14], while the measured values of the radius of
the curvature for the micropillars are very small
(Fig. 7b), most of which are dispersed in the range from
2 to 9 nm.
Eq. (3) was rst used to extract the pillars Youngs
modulus in the same way as in the literature [2,5,6,15,16].

Fig. 5. The representative geometries of the FIB-fabricated micropillars


with markings on the pillars outline showing how their geometrical
features were measured: the blue arrows are for the top diameter; the red
arrows for the bottom diameter and the green dashed lines for the taper
angle (note that the vertical length scale diers from the horizontal one
due to the sample stage tilting). (For interpretation of color mentioned in
this gure the reader is referred to the web version of the article.)

4. Results and discussions


Prior to the microcompression tests, quantitative
assessments of the pillar geometries were attempted.

Fig. 6. Symbols used in the derivation of the radius of the curvature at the
base of the micropillar.

Fig. 7. Bar charts of the measured values for the taper angle (a) and the
radius of the curvature (b) for the Zr50Al10Cu37Pd3 micropillars.

1618

Y. Yang et al. / Acta Materialia 57 (2009) 16131623

Fig. 8. Comparison of the Youngs moduli obtained from the microcompression tests before and after correction.

For the micropillars of approximately the same aspect


ratios, similar results were obtained for their Youngs moduli. However, when putting all the measurements together,
a spurious geometry dependence of the micropillars
Youngs modulus on its aspect ratio became visible
(Fig. 8). By averaging out the experimental data of all
the micropillars, the measured Youngs modulus was about
76 12 GPa.
In contrast, after the correction was made in accord with
Eq. (5) and Fig. 3, for which the Youngs modulus previously obtained for each micropillar was multiplied by a
pre-factor, w, that relates to the pillar geometry, a geometry-independent value for the pillars Youngs modulus was
extracted as shown in Fig. 8, which was about 99 6 GPa.
Compared with the scatter in the uncorrected measurements, the scatter in the corrected measurements is now
reduced from 16% to 6%, indicative of the improved
repeatability after the correction.
For the sake of comparison, nanoindentation experiments were conducted on the surface of the same bulkmetallic glass sample using a Berkovich indenter to extract
the Youngs modulus. Only the forcedisplacement curves
that followed a master loading curve were selected in the
data analysis, and the results were plotted in Fig. 9. Given
the Poisson ratio of 0.366 for the Zr-based bulk-metallic
glass [24], the Youngs modulus of 129 6 GPa was
obtained, which is about 30% higher than that from the
microcompression tests. As pointed out in Refs. [25,26],
there is a tendency for the nanoindentation tests, which
are based on the Oliver and Pharr method [11], to overestimate the Youngs modulus of bulk-metallic glasses. This
is because the plastic ow in bulk-metallic glasses, having
nearly zero strain hardening, is prone to a pile-up around
an indent [27], which leads to the underestimation of the
indentation area using the Oliver and Pharr method
(Fig. 10). In order to compensate for the pile-up eect,
an improved indentation method taking into account the
materials pile-up was used to correct the measured

Youngs moduli2 [28], after which the measured Youngs


modulus was reduced to 103 4 GPa. In addition,
atomic-force microscopy (AFM) was employed to image
the indent proles for the extraction of the accurate indentation contact area for further verication. Following the
method in Ref. [28], the true indentation contact area
was taken as the area encircled by the pile-up material
as apposed to the triangular area related to the indentation
contact depth through the geometrical similarity, which
yielded results similar to those from the improved indentation method (Fig. 11).
Table 1 lists the measurements of the Youngs modulus
of the same Zr-based metallic glass from the three dierent
methods. It can be seen that both the microcompression
and nanoindentation tests yielded very similar results.
Although the Youngs modulus obtained from the conventional compression tests is very close to the lower bound of
the measurements from the microcompression tests, it is,
however, about 10% lower than the average values
obtained from both the microcompression and nanoindentation tests. Although it is still not very convincing about
this discrepancy since no experimental scatter was provided
for the Youngs modulus obtained from the bulk samples
in the literature [24], it is not uncommon that the pre-existing casting defects will reduce the eective Youngs modulus when a large sample is tested. Given the large dierence
in the length scale of the sampling volume among the three
methods (Table 1), the 10% dierence in the measurements
is regarded as acceptable.
Before closing this section, it is worthwhile having a brief
discussion on the other factors that may aect the microcompression tests of metallic glasses. When the aspect ratio of the
micropillar is over 10, it is prone to elastic buckling due to
the misalignment between the indenter and micropillar.
Consequently, a quite dierent forcedisplacement curve
was acquired as compared to those from the non-buckling
tests, giving rise to a Youngs modulus measurement much
lower than otherwise (Fig. 12). Given the dramatic change
in the forcedisplacement curve, the experimental data corresponding to elastic buckling can be easily excluded in the
subsequent data analysis. On the other hand, the elastic
and geometrical mismatch between the indenter and the
micropillar causes a stress concentration at the lateral
boundary of the indenter/micropillar contact. In conventional compression tests on BMGs [2933], such a kind of
stress concentration at the interface between a platen and a
compression specimen leads to an aspect ratio eect on the
compressive behavior of the BMG specimens. In the microcompression experiments, the aforementioned stress concentration still exists and partly contributes to the stress nonuniformity in the micropillars. In the elastic deformation
regime, the stress concentration eect was captured by the
nite-element simulations for the case of non-frictional
2

The initial yield strength of 1899 MPa and zero strain hardening
exponent, obtained from the macroscale tensile experiments [28], were
used in the correction.

Y. Yang et al. / Acta Materialia 57 (2009) 16131623

1619

Fig. 9. Variation of the metallic glass Youngs modulus with the indentation contact depth (the inset to the left shows the AFM height image of an
indentation impression and the inset to the right shows the loaddisplacement curves).

Fig. 10. The 3-D rendered AFM image of an indentation impression


made by a Berkovich indenter showing the pile-up around the periphery.

indenter/micropillar contact; while in the plastic deformation regime, it is still an unexplored issue to date whether
the stress concentration can lead to such an aspect ratio eect
on the post-yielding behavior of the metallic-glass micropillars. As for the yielding point of metallic-glass micropillars, it
was found that, due to the stress concentration, the measurement of the yielding strength was sensitive to the constitutive
behavior assumed, and interested readers are referred to the
related work by Schuster et al. [34] for a detailed discussion.
5. Implications
Based on the above analyses and discussions, we have
demonstrated that, as long as the geometrical features of

Fig. 11. The true indentation area taking into account the pile-up, as
outlined by the dashed line, vs. the computed indentation area as outlined
by the dotted line in a 2-D AFM image.

the micropillar are measured from the SEM pictures, the


microcompression tests are capable of measuring the elastic
modulus of the metallic glass with a similar accuracy as the
nanoindentation tests. However, in view of the relatively
simple structure of the testing specimen at such a small
scale, a few applications could be envisioned for the microcompression tests.
Over the past decades, great eorts have been dedicated
to the development of a reliable method to measure the
mechanical properties of a coating/substrate system, which
represents a class of industrial and academic problems of
great importance. In the past, the focus has been directed
to the use of nanoindentation to study coating/substrate
systems [35]. However, the coating/substrate interactions

1620

Y. Yang et al. / Acta Materialia 57 (2009) 16131623

Table 1
Summary of the Youngs modulus measurements from the three dierent
methods.
Quantity method

Youngs modulus
(GPA)

Length-scale of the
sampling volume

Macrocompression tests [24]


Microcompression tests
Nanoindentation tests

90
99 6
103 4

>1 mm
0.53 lm
1.59 lma

Note that the length scale of the sampling volume in a nanoindentation


test is taken as the plastic zone size, c, approximated by the Johns
indentation model [49] as c = (9P/2pH)0.5, where P is the indentation load
and H is the corresponding indentation hardness.

complicate the mechanical responses and the interpretation


of the experimental data is therefore non-trivial, which
relies on the prior knowledge of a few key properties that
are related to materials heterogeneity around the interface
between the coating and substrate [36]. On the other hand,
the microcompression methodology may provide an alternative solution to the aforementioned problems by separating the coated material from the substrate in a single test.
Since the deformation is limited to the micropillar, the concerns related to the interactions between dierent constituents in a multi-structured/multi-phased system can, to
some extent, be relieved. However, given the assumptions
of material homogeneity and isotropy, our method is
now only applicable to a specic type of coating/substrate
systems. One recent novel application of metallic glasses is
to sputter a metallic-glass thin lm onto the surface of a
classic engineering material to enhance the fatigue resistance as shown in Fig. 13a [3740]. In these applications,
it is extremely dicult to obtain a reliable assessment of
the mechanical properties of the coating using the nanoindentation method as the confounded mechanical
response is site-sensitive due to the material heterogeneity
in the underlying structural materials. Now, one plausible

Fig. 13. FIB images of the complex materials systems suitable for
microcompression tests for the measurement of the mechanical properties
of an individual constituent: (a) a metallic glass coating on a stainless steel
substrate and (b) a nanocrystalline nickel iron alloy with a grain size
gradient (the red arrows indicate the cross-section/top surface transition).
(For interpretation of color mentioned in this gure the reader is referred
to the web version of the article.)

solution to overcome such a confounding eect from the


substrate is to completely separate the thin lm from the
substrate by milling out a small-scale specimen in the area
of interest for the microcompression test.
The other type of material systems that attracts the interest of many materials scientists and is suitable for microcompression study now is the so-called plasticity gradient
materials. One method to realize the plasticity gradient is
to introduce a grain size gradient using the well-established

Fig. 12. Comparison of the buckling vs. non-buckling forcedisplacement data acquired from the microcompression tests.

Y. Yang et al. / Acta Materialia 57 (2009) 16131623

surface-mechanical-attrition-treatment (SMAT) technique


[4145]. After SMAT, a grain size distribution, similar to
that shown in Fig. 13b, can be implemented into a polycrystalline metal for the optimization of its overall mechanical
properties. In spite of the overall heterogeneity, some local
homogeneity and isotropy still remain in the areas of similar
grain size. Therefore, the local mechanical properties can be
measured using the micropillars as long as they sample a
high enough number of similar-sized grains to ensure the
homogeneity and isotropy.
6. Summary and conclusions
In summary, we have demonstrated that the approximation with the Sneddons formula is insucient to obtain an
accurate measurement of the pillars Youngs modulus.
Based on the simulations and experiments, an improved
formula can herein be proposed, which is:


p1  mD0
4PH
E W 1
8H
pD0 D0 2H tanhdt

where w is the geometry-dependent pre-factor and its value


can be found in Fig. 3ad according to the aspect ratio, the
taper angle and the radius of the curvature of the micro/
nanopillar concerned, which are all measurable quantities
in a microcompression experiment. Since the taper angle
for the appropriately ion-beam-treated micro/nanopillar
side walls ranges from 0 to 6, the aspect ratio mostly used
in the literature is around 23 and the radii of the curvature
measured are all within 10% of the corresponding top radius, we would expect that Fig. 3ad has comprised most
of the w values that might be used for future microcompression studies.
The advantage of a microcompression test over a nanoindentation test is the relatively simple stress state that is
limited within the micropillar. As long as the micropillar
is placed far away from the boundary susceptible to a

microstructural change, Eq. (7) is applicable and a reliable


measurement of the elastic properties should be warranted.
Acknowledgments
Y. Y acknowledges the internal research funding support to this work granted by the Hong Kong Polytechnic
University through the Department of Mechanical Engineering. The bulk-metallic glass used in the present work
is provided by Prof. Y. Yokoyama of Tohoku University
and Dr. G.Y. Wang of The University of Tennessee.
F.X.L. and P.K.L. are grateful to the support of the National Science Foundation, International Materials Institutes (IMI) Program (DMR-0231320) with Dr C. Huber
as the Program Director.
Appendix A
In conventional nanoindentation tests, the machine
compliance is usually calibrated prior to data acquisition
with its eect being eliminated from the subsequently
acquired displacement data. However, after many years
of use, the machine compliance may deviate from the
pre-dened or built-in value and, therefore, needs to be
re-calibrated. In the case of an unknown machine compliance Cf, the measured total displacement, dt, is the sum
of the displacements in the material, dm, and the loading
frame, which is:
dt dm C f P

A1

To our knowledge, there are two methods available in


the literature to calibrate the machine compliance of a nanoindentation system [11,46]. For brevity, only the method
proposed by Sun et al. [46] is introduced below for the
study of our nanoindentation system. Since the shape of
the diamond nanoindenter only possesses a negligible eect
on the loading frame compliance, a conventional Berkovich nanoindenter can be used to estimate Cf. Without
prior knowledge of the loading frame compliance Cf and
the indenter tip radius r in a real indentation experiment
of isotropic and homogeneous materials, the total indentation displacement h can be expressed as:
h C f P K 1=2 P 1=2  g

Fig. A1. The plot of one of the h  P1/2 curves used for the calibration of
the machine compliance with thermal drift corrected.

1621

A2

where K is a constant that depends on the materials


mechanical properties and g is a geometrical parameter
that is related to the indenter tip radius and thermal drift.
Fitting the Berkovich nanoindentation data obtained on
the Zr50Al10Cu37Pd3 BMG sample with Eq. (A2), the
machine compliance was extracted with its value ranging
from 0.0001 to 0.0007 nm mN1 (one of the extractions is
shown in Fig. A1), which conrmed that the eect of the
machine compliance had been ruled out in our experiments
with a pre-dened machine compliance of 0.5 nm mN1.
Meanwhile, the g value of 15.3 nm corresponds to a tip
radius of 230 nm for the nanoindenter, which is reasonable for a Berkovich-type indenter already tip-rounded

1622

Y. Yang et al. / Acta Materialia 57 (2009) 16131623

tapered ends. The critical buckling load, Pcr, for such a


straight bar, with the two ends xed and subjected to a uniaxial point load, is given by [47]:
P cr

Fig. A2. The 2-D schematics of (a) the original micropillarsubstrate


system, (b) the double tapered straight bar with a uniform mid-section for
the upper bound estimate of the critical buckling load, and (c) the overconstrained micropillar for the lower bound estimate of the critical
buckling load.

UEI b
H2

A3

where H is the micropillars height, E the micropillars


Youngs modulus, Ib is the second moment of inertia for
the bottom surface of the micropillar and U is the pre-factor that depends on the geometrical parameters, a/L, and
It/Ib (It = the second moment of inertia for the top surface
of the micropillar) as shown in Fig. A2b, whose values were
obtained through numerical analysis and tabulated in Ref.
[47]. For the system of the metallic-glass micropillars, a/
L  1 and It/Ib varies between 1.2 and 1.6. The linear
extrapolation of the available data in Ref. [47] gives an
average value of the pre-factor U as 42.
The critical aspect ratio is attained with the critical
buckling load Pcr equal to the critical load PY at yielding.
For the slightly tapered micropillars, assuming
P Y pD20 rY =4 gives:
s
H
1 10E

A4
D0 2 rY

after repeated use (the original tip radius was 150 nm).
This implies that the thermal drift was also appropriately
corrected in the microcompression experiments. Based on
the above discussions, the use of Eq. (1) is justied for data
analysis.

where D0 is the diameter of the micropillars top surface.


For the Zr-based micropillars, their yield strength is
roughly equal to that of their bulk counterparts. Substituting the measured values E = 90 GPa and rY = 1.9 GPa
into Eq. (A4) yields the upper bound estimate of 10.88
for the critical aspect ratio.
In contrast, imposing the xed boundary conditions at
both the top and base of the micropillar, as shown in
Fig. 2Ac, corresponds to the lower bound estimation of
the critical buckling load [48]:

Appendix B

P cr

Due to the complexities in the geometry of the micropillarsubstrate system, it is non-trivial to obtain a close-form
solution to the critical buckling load. Instead, we are trying
to seek its upper and lower bound estimates by loosening
and strengthening the mechanical constraints from the substrate, respectively.
To make an upper bound estimate, let us assume that the
constraint from the substrate on the micropillar is restricted
to the materials in a cylindrical volume right below the
micropillars base (as shown by the hatched area in
Fig. A2a). This cylindrical substrate possesses the height
of the original substrate, denoted as a/2 in Fig. A2b, and
the same diameter, Db, as that of the micropillars base. Since
H/a<<1, the remote boundary of the cylindrical substrate is
unaected by the stresses in the micropillar and regarded as
xed. Due to the symmetry, the critical buckling load of the
micropillarcylindrical-substrate system is identical to that
of a straight bar with the two end portions tapered and the
mid-section uniform (Fig. A2b). Owing to the tapering in
the end portions, the rst buckling mode occurs in the

4Xp2 EI 0
A5
H2
For the slightly tapered FIB-milled micropillars, the corresponding pre-factor X is 1 [48]. Repeating the similar
procedure as above gives the lower bound estimate of the
critical aspect ratio:
r
H
p E
A6

D0 2 rY
Similarly, the lower bound estimate of the critical aspect
ratio can be obtained as about 10.81 by substituting
E = 90 GPa and rY = 1.9 GPa into Eq. (A6).
In summary, the above upper and lower bound analyses
indicate that the actual value of the critical aspect ratio for
our micropillars should be 11, which agrees with our experimental observations that no elastic buckling occurred for
the micropillar with an aspect ratio smaller than 10.
References
[1] Dimiduk DM, Uchic MD, Parthasarathy TA. Acta Mater 2005:53.
[2] Greer JR, Oliver CW, Nix WD. Acta Mater 2005:53.

Y. Yang et al. / Acta Materialia 57 (2009) 16131623


[3] Moser B, Wasmer K, Barbieri L, Michler J. J Mater Res 2006;22:1004.
[4] Bei H, Shim S, George EP, Miller MK, Herbert EG, Pharr GM, et al.
Scripta Mater 2007;57:397.
[5] Lee CJ, Huang JC, Nieh TG. Appl Phys Lett 2007;91:161913.
[6] Lai YH, Lee CJ, Cheng YT, Chou HS, Chen HM, Du XH, et al.
Scripta Mater 2008;58:890.
[7] Zheng Q, Cheng S, Strader JH, Ma E, Xu J. Scripta Mater
2007;56:161.
[8] Shan ZW, Li J, cheng YQ, Minor AM, Syed Asif SA, Warren OL,
et al. Phys Rev B 2008;77:155419.
[9] Donohue A, Spaepen F, Hoagland RG, Misra A. Appl Phys Lett
2007;91:241905.
[10] Cheng S, Wang XL, Choo H, Liaw PK. Appl Phys Lett
2007;91:201917.
[11] Oliver WC, Pharr GM. J Mater Res 1992;7:1564.
[12] Juan JMS, No ML, Schuh CA. Adv Mater 2008;20:272.
[13] Volkert CA, Donohue A, Spaepen F. J Appl Phys 2008;103:083539.
[14] Volkert CA, Lilleodden ET. Philos Mag 2006;86:5567.
[15] Zhang H, Schuster BE, Wei Q, Ramesh KT. Scripta Mater
2005;54:181.
[16] Choi YS, Uchic MD, Parthasarathy TA, Dimiduk DM. Scripta
Mater 2007;57:849.
[17] Sneddon IN. Int J Engrg Sci 1965;3:47.
[18] Wang GY, Liaw PK, Yokoyama Y, Peker A, Peter WH, Yang B,
et al. Intermetallics 2007;15:663.
[19] Gu XJ, Poon SJ, Shiet GJ. J Mater Res 2007;22:344.
[20] Zhang Y. Mater Sci Technol 2008;24:379.
[21] Schroers J, Johnson WL. Phys Rev Lett 2004;93:255506.
[22] Lewandowski JJ, Wang WH, Greer AL. Philos Mag Lett 2005;85:77.
[23] Poon SJ, Zhu A, Shiet GJ. Appl Phys Lett 2008:92.
[24] Wang GY, Liaw PK, Yokoyama Y, Inoue A, Liu CT. Mater Sci Eng
A 2008. doi:10.1016/j.msea.2008.04.03.
[25] Shu SQ, Lu J, Li DF. J Mater Res 2007;22:3385.
[26] Wright WJ, Saha R, Nix WD. Mater Trans 2001;424:642.
[27] Cheng YT, Cheng CM. Philos Mag Lett 1998;78:115.

[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]

1623

Choi Y, Lee HS, Kwon D. J Mater Res 2004;19:3307.


Sunny G, Yuan F, Prakash V, Lewandowski JJ. Exp Mech 2008.
Sunny G, Lewandowski JJ, Prakash V. J Mater Res 2007;22:389.
Jiang WH, Fan GJ, Choo H, Liaw PK. Mater Lett 2006;60:3537.
Liu FX, Liaw PK, Wang GY, Chiang CL, Smith DA, Rack PD, et al.
Intermetallics 2006;14:1014.
Zhang ZF, Zhang H, Pan XF, Das J, Eckert J. Philos Mag Lett
2005;85:513.
Schuster BE, Wei Q, Hufnagel TC, Ramesh KT. Acta Mater
2008;56:5091.
Cheng X, Vlassak JJ. J Mater Res 2001;16:2974.
Gerberich WW, Kramer DE, Tymiak NI, Volinsky AA, Bahr DF,
Kriese MD, et al. Acta Mater 1999;47:4115.
Liu FX, Chiang CL, Chu JP, Gao YF, Liaw PK. Mater Res Sco
Symp Proc 2006;903E:0903.
Liu FX, Liaw PK, Jiang WH, Chiang CL, Gao YF, Guan YF, et al.
Mater Sci Eng A 2007;468:246.
Liu FX, Gao YF, Liaw PK. Metall Mater Trans A 2008;39A:1862.
Chiang CL, Chu JP, Liu FX, Liaw PK, Buchanan RA. Appl Phys
Lett 2006;88:131902.
Zhang HW, Hei ZK, Liu G, Lu J, Lu K. Acta Mater 2003;51:1871.
Ortiz AL, Tian JW, Villegas JC, Shaw LL, Liaw PK. Acta Mater
2008;56:413.
Villegas JC, Dai K, Shaw LL, Liaw PK. Mater Sci Eng A
2005;410:257.
Villegas JC, Shaw LL, Dai K, Yuan W, Tian J, Liaw PK, et al. Philos
Mag Lett 2005;85:427.
Tian JW, Villegas JC, Yuan W, Fielden D, Shaw LL, Liaw PK, et al.
Mater Sci Eng A 2007;468:164.
Sun Y, Zheng S, Bell T, Smith J. Philos Mag Lett 1999;79:649.
Young WC, Budynas RG. Elastic stability. Roaks formulas for stress
and strain. New York: McGraw-Hill; 2002. p. 725.
Gere JM, Carter WO. American Society of Civil Engineers Transactions 1963;128:736.
Choi Y, Suresh S. Scripta Mater 2003;48:249.

Vous aimerez peut-être aussi