Sunteți pe pagina 1din 13

Acta Materialia 55 (2007) 30593071

www.elsevier.com/locate/actamat

Temperature, strain rate and reinforcement volume fraction


dependence of plastic deformation in metallic glass matrix composites
X.L. Fu a, Y. Li
b

a,b

, C.A. Schuh

c,*

a
Singapore-MIT Alliance, National University of Singapore, Lower Kent Ridge Road, Singapore 119260, Singapore
Department of Materials Science and Engineering, Faculty of Engineering, Engineering Drive 1, National University of Singapore,
Singapore 117576, Singapore
c
Department of Materials Science and Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139, USA

Received 15 November 2006; received in revised form 9 January 2007; accepted 14 January 2007
Available online 2 March 2007

Abstract
A systematic study of mechanical properties is presented for Zr-based bulk metallic glass matrix composites, spanning a wide range of
strain rates and temperatures, as well as various levels of reinforcement volume fraction. All of the experimental materials exhibit mechanical properties dominated by deformation of the amorphous matrix phase, including inhomogeneous ow and fracture at low temperatures, as well as homogeneous ow of both Newtonian and non-Newtonian character at high temperatures. In the homogeneous ow
regime, the composites exhibit clear strengthening as the volume fraction of reinforcement increases. This strengthening eect is quantitatively explained in both the Newtonian and non-Newtonian regimes, and is found to arise from two contributions: (i) load transfer from
the amorphous matrix to the reinforcements; and (ii) a shift in the glass structure and properties upon precipitation of the reinforcements.
An additional source of apparent strengthening in situ precipitation of reinforcement during deformation is also discussed.
 2007 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Keywords: Metallic glass; Composites; Creep; Mechanical properties; High temperature deformation

1. Introduction
The unique properties of bulk metallic glasses (BMGs)
include generally high yield strengths of up to several
GPa, elastic strain limits of 2% at room temperature, generally excellent wear and corrosion resistance, as well as
good shaping and forming ability in the viscous state [1
4]. However, most BMGs exhibit very little microscopic
plasticity when deformed at room temperature, as they
tend to fail catastrophically along one dominant shear
band. During the past 10 years, some enhancement of the
macroscopic plasticity of BMGs has been gained by adding
secondary crystalline phases into the bulk metallic glass
matrix, to form bulk metallic glass matrix composites
(BMGMCs) [514]. Reinforcing bers, ceramics and duc-

Corresponding author.
E-mail address: schuh@mit.edu (C.A. Schuh).

tile particles have all been explored in this context. For


example, Mo, Nb and Ta [9] particles have been explored
as reinforcements to Zr-based BMGs to improve tensile
ductility to values as high as 17%. In 2000, Hays and Johnson [15] introduced a new class of BMGMC, synthesized
in situ during cooling from the melt, whereupon a ductile
dendritic phase precipitated in the metallic glass matrix.
Those authors demonstrated improved ductility from near
zero (for the base matrix material) to around 5% (for the
composite), and other researchers have subsequently studied in situ BMGMC processing in various alloy systems.
Lee et al. [13] demonstrated the ability to tailor the volume
fraction of in situ reinforcements in a family of La-based
BMGMCs, and explored the role of volume fraction on
strength and ductility at room temperature.
Aside from composite reinforcements, another
approach to increase the ductility and formability of BMGs
is to induce homogeneous ow at elevated temperatures and
low strain rates [1624]. Whereas low temperatures (usually

1359-6454/$30.00  2007 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.actamat.2007.01.009

3060

X.L. Fu et al. / Acta Materialia 55 (2007) 30593071

including room temperature) and/or high strain rates generally promote intense strain localization into shear bands,
increasing the temperature to around 70100% of Tg (the
glass transition temperature) and decreasing the strain rate
to relatively low values leads to homogeneous, viscous
deformation in BMGs; in many cases superplastic-like ow
is exhibited under these conditions [25]. Details of the
homogeneous ow of metallic glasses were rst described
in the 1970s [19,2631], and more recent studies by Nieh
and Wadsworth [32], Spaepen [33] and others [3437] have
investigated this issue further. Lu et al. [38] made a general
investigation of stressstrain relations for ZrTiCuNi
Be glass over a broad range of temperatures and strain
rates.
Given that macroscopic ductility is promoted in BMGs
by both composite reinforcements and high-temperature
viscous ow, it is of fundamental interest to understand
deformation physics when both of these strategies are
employed at the same time. However, despite the quickly
growing literature around the improved room temperature
mechanical properties of reinforced metallic glasses, the
deformation behavior of BMGMCs over a wide range of
temperatures and strain rates has yet to be systematically
investigated, especially at high temperatures close to Tg.
The work of Bae et al. presented some tensile stressstrain
data for ex situ Ni-based BMGMCs [39] and an in situ Tibased BMGMC [40] tested in the supercooled liquid
region, but did not identify the impact of the reinforcement
on the material rheology. In a recent preliminary report, we
presented data concerning the strain-rate sensitivity of

homogeneous ow in Zr-based BMGMCs [41], but only


at a single test temperature near Tg.
It is our purpose in what follows to elucidate the eects
of temperature, strain rate and reinforcement volume fraction upon the plastic deformation of in situ Zr-based
BMGMCs. This work signicantly expands upon our preliminary study of the same materials [41], and focuses upon
the regime of homogeneous ow and the transition to this
mode of deformation at low rates and high temperatures.
Results for three dierent volume fractions of in situ reinforcement are compared with the deformation behavior of
the fully amorphous matrix, and rationalized on the basis
of rheological models.
2. Experimental
The base composition of the materials used in this work
was Zr49Cu(51x)Alx (x = 6, 8, 10 or 12 at.%), prepared by
arc-melting a mixture of Zr (99.9% purity), Cu (99.99%)
and Al (99.9%) in an argon atmosphere. The arc-melted
liquid metal was subsequently cast into a copper mold to
produce rods 5 mm in diameter and around 56 mm in
length. The four dierent values of aluminum content (x)
investigated led to four microstructures that developed
in situ upon casting, as depicted in the scanning electron
microscope (SEM) images in Fig. 1. In these images the
glassy matrix appears darker, while a second dendritic
phase appears lighter. By using the commercial software
Leica Qwin Image Analyzer, the volume fraction of these
reinforcing dendrites (denoted as f) was found to increase

Fig. 1. Scanning electron micrographs of alloys with composition Zr51Cu49xAlx where x = 6, 8, 10, and 12 at.%, in (ad), respectively. The volume
fraction of second-phase dendrites is (a) 0%, (b) 7%, (c) 15% and (d) 20%. Adapted from Ref. [41].

X.L. Fu et al. / Acta Materialia 55 (2007) 30593071

monotonically with aluminum content (f = 0%, 7%, 15%


and 20% for values of x = 6, 8, 10 and 12 at.%, respectively). In what follows we refer to the four dierent alloys
by their respective values of f, rather than their
compositions.
Mounted and polished cross-sections of the as-cast rods
were examined using conventional X-ray diractometry
(XRD, using a Bruker analytical X-ray system), to identify
the phases present. Additionally, the glass transition as
well as the crystallization temperature were characterized
by dierential scanning calorimetry (DSC, using a TA
Instruments 2920 modulated DSC) at a heating rate of
20 K min1. Specimens of each composition were also subjected to standard compression testing, using cylindrical
specimens cut from the cast rods with lengths of about
10 mm and diameters of about 5 mm. Each compression
specimen was ground parallel to an accuracy generally better than 20 lm, using SiC grinding papers, followed by polishing with 0.05 lm alumina powder; the lateral surfaces of
the cylinder were also carefully polished through a similar
regimen.
The compression tests were conducted at various temperatures, using a three-zone clamshell furnace with temperature control to within 2 K, mounted on an Instron
4505 machine. The displacement rate of the crosshead
was held constant for any given test, and was varied to give
engineering strain rates on the range 1 105 to
1.7 102 s1. The compression platens were composed
of a Ni-based superalloy (MAR-M 246), in a compression
cage conguration that guaranteed axial alignment. The
experiments were performed at various constant temperatures ranging from 463 to 715 K. Specimens were always
loaded into the furnace at the intended test temperature,
and monitored by a K-type thermocouple placed about
1 cm away from the specimen in the center of the furnace.
Generally about 20 min was required to equilibrate the system, after which the test was initiated. After each compression test, the sample was immediately taken from the
furnace and air cooled to room temperature. Depending
upon the strain rate employed, the samples generally experienced between 25 and 230 min of elevated temperature
exposure. In total, 75 dierent experiments were conducted
at various combinations of applied strain rate, temperature
and alloy composition.
For all of the mechanical tests, the stress state was monitored through the use of a calibrated load cell, and the
length of the specimen was measured using conventional
high-temperature extensometry and a linear voltagedisplacement transducer. All of the compression test data presented here are given in terms of true stresses and strains.

3061

The f = 0% alloy is the baseline bulk metallic glass-forming


alloy [42], and was conrmed by XRD to be amorphous
(Fig. 2). In contrast, the other three compositions with
higher aluminum content exhibit dispersed second-phase
particulates, mostly in the form of ne dendrites. In all
three of these composite-forming compositions, the dendritic phase is believed to be the so-called s3 phase,
Zr51Cu28Al21 [42], based on the position of crystalline
peaks seen in the XRD patterns of Fig. 2; these reinforcements are generally on the order of 510 lm in breadth. At
the lower values of f (=715%), the dendritic phases appear
to have a characteristic fourfold symmetry, but in the
f = 20% alloy there are also some larger, more complex
rosette morphology reinforcements, apparently of the same
s3 phase.
Fig. 3a and Table 1 show the onset T onset
and the end
g
end
T g of the endothermic glass transition event, as well as
the crystallization temperature (Tx) of each specimen. It is
interesting to observe that changing the alloy composition
not only promotes the formation of in situ s3 reinforcements in this system, it also somewhat aects the composition of the glassy matrix. This result is reected by a shift in
the values of Tg with composition, which is plotted in
Fig. 3b for T onset
. Table 1 also lists the total exothermic
g
heat released during crystallization of each alloy, which
monotonically decreases as the second-phase volume fraction increases; due to the lower fraction of the amorphous
matrix in the composite specimens, less heat is involved
upon crystallization of the matrix.
3.2. Compressive deformation behavior
Fig. 4 shows a representative series of true stresstrue
strain curves measured in compression for the four experimental alloys at a constant test temperature of 693 K, and
a variety of applied strain rates. Several dierent curves
have been presented on each set of axes, oset from one
another in strain for clarity. Depending upon the strain

3. Results
3.1. Materials characterization
As introduced earlier, the SEM micrographs of Fig. 1
show the microstructures of the four experimental alloys.

Fig. 2. X-ray diraction patterns of the base amorphous specimen


(f = 0%) and composite samples with increasing volume fractions of
dendritic second phase, indexed here as the s3 phase.

3062

X.L. Fu et al. / Acta Materialia 55 (2007) 30593071

Fig. 3. (a) DSC scans for the monolithic metallic glass Zr49Cu51xAlx (with x = 6) as well as in situ composites with dierent aluminum contents (x = 8,
10, 12); the specimens are labeled with their respective second-phase volume fractions, f, which increase with aluminum content. (b) The onset glass
transition temperature is plotted for all of the as-cast samples, as a function of the volume fraction of second phase. Also shown are experimentally derived
T onset
values for specimens deformed at 693 K, which are close to the values for the as-cast specimens. Part (a) is adapted from, and some of the data in (b)
g
appeared previously in Ref. [41].

Table 1
and end T end
Characteristic temperatures describing the onset T onset
g
g of
the glass transition, as well as the crystallization temperature (Tx), width
of the supercooled liquid region (DTx) and heat release of crystallization
(DHx) obtained from DSC measurements of the as-cast samples
Zr sample

T onset
K
g

T end
K
g

Tx (K)

DTx (K)

DHx (J g1)

f = 0%
f = 7%
f = 15%
f = 20%

694
699
704
707

709
712
720
722

755
761
763
769

61
62
59
62

58
57
46
36

rate applied, both inhomogeneous and homogeneous


deformation are observed. Low strain rates generally promote homogeneous ow with a pronounced steady-state
regime at a constant true stress. At somewhat higher rates
a characteristic stress overshoot evolves before the system
reaches a steady state, and at the highest rates, apparently
brittle failure occurs via inhomogeneous shear fracture. All
of the specimens, including all three composite volume
fractions, exhibited the stress overshoot phenomenon at
certain strain rates.
Fig. 5 shows a complementary set of representative
stressstrain curves, now all at a constant applied strain
rate of 1.7 103 s1 but at various test temperatures.
Below about 663 K, the specimens all exhibit the same
characteristic behavior, with elastic deformation followed
by inhomogeneous, catastrophic shear failure. At intermediate temperatures, up to about 693 K, the composite specimens exhibit an initial stress overshoot followed by an
extended regime of homogeneous, steady-state ow, while
the amorphous alloy more quickly transitions to a steady
state condition without any overshoot stress. At even
higher temperatures, above about 703 K, the elastic regime
transitions directly into a steady-state ow curve (without
overshoot) even for the composites.

Figs. 6 and 7 summarize the basic trends that we have


measured in the mechanical properties of BMGMCs. In
Fig. 6, we plot the characteristic stresses measured in each
experiment as a function of the applied strain rate at 693 K.
Fig. 6a shows the eect of rate on the peak stress measured,
which can represent either the fracture stress or the overshoot peak stress; for tests in which no overshoot appears
(i.e. steady state is established directly upon loading) we
have not presented any data in Fig. 6a. In general, the rate
dependence of the peak stress is weak, especially at higher
rates, where homogeneous ow is less likely. In contrast,
Fig. 6b plots the steady-state ow stress, and a stronger
rate dependence is discerned (only data from tests that
exhibited a steady-state regime are included in Fig. 6b).
In both of these cases, alloy composition/structure has a
signicant eect on the mechanical response, with higher
reinforcement loading yielding palpably higher strength.
Fig. 7a and b shows the complementary data for peak
and ow stress as a function of temperature for a constant
rate. At low temperatures, where inhomogeneous deformation prevails, the general trend of the peak stresses is not
strongly temperature-dependent. At roughly 625 K,
strength drops quickly owing to the transition to homogeneous ow. In the high-temperature regime, there is clear
evidence for an increase in ow stress with reinforcement
loading (Fig. 7b).
It is interesting to note the variability in the strength
data of each individual material at lower temperatures.
There is no clear trend with respect to temperature; in fact,
there is an apparent peak in strength in some of the data
sets in Fig. 7a at an intermediate temperature close to the
high-temperature transition point. These observations all
speak to the natural variability of strength in the inhomogeneous regime, where the material is quite defect-sensitive,
and failure occurs rapidly when the yield condition is met
anywhere along the length of the specimen. Sensitivity to

X.L. Fu et al. / Acta Materialia 55 (2007) 30593071

3063

Fig. 4. True stressstrain curves for the experimental materials with (a) f = 0%, (b) f = 7%, (c) f = 15% and (d) f = 20% second-phase reinforcements.
These graphs specically illustrate the eect of applied strain rate on the uniaxial compression response at T = 693 K.

Fig. 5. True stressstrain curves highlighting the eect of temperature on the uniaxial compressive behavior of the monolithic amorphous alloy (a) and its
composites (bd) at a constant imposed strain rate 1.7 103 s1.

3064

X.L. Fu et al. / Acta Materialia 55 (2007) 30593071

Fig. 6. Summary of the mechanical test data acquired at T = 693 K, illustrating rate eects on characteristic strengths. (a) Peak stress is plotted as a
function of strain rates for the monolithic amorphous alloy and its composites, for specimens which either exhibited a stress overshoot or fractured prior to
measurable plastic deformation. (b) Steady-state ow stress is plotted against applied strain rate on double-logarithmic scales. In (a) the lines are presented
only for visualization purposes, while in (b) the lines are best-t forms of Eq. (1). Data in (b) appeared previously in Ref. [41].

Fig. 7. Summary of the mechanical test data acquired at a constant applied strain rate of 1.7 103 s1, illustrating the eect of test temperature on
strength. In (a) peak stress, and in (b) the steady-state ow stress are plotted as a function of temperature. The lines in (a) are presented to illustrate trends
only, while in (b) they represent the best ts of Eqs. (1) and (2).

casting defects, microstructure inhomogeneity and other


stress concentrators can lead to such scatter in strength
data. Furthermore, as the temperature is increased to a sufciently high level approaching the transition to homogeneous ow, the material undergoes a transition to a state
in which it resists catastrophic shear failure and is no
longer defect-sensitive. Whereas at low temperatures
defect-sensitivity can lead to articially low fracture
strength measurements, the behavior is more deterministic
at higher temperatures; the peak strength seen at intermediate temperatures is probably a consequence of this
cross-over.
3.3. Microstructural evolution
Deformed samples of each experimental material are
compared with as-cast samples in the XRD patterns of
Fig. 8. Alongside each curve the conditions under which

the specimen was tested (temperature, rate and total


time-at-temperature) are also shown. Referring rst to
the data in Fig. 8a and b, for the base glass composition
(f = 0%) and the most dilute composite system studied
(f = 7%), we see no remarkable changes in the diraction
patterns after testing. However, the more concentrated
composites, f = 15% and 20% in Fig. 8c and d, show clear
signs of structural evolution during the high-temperature
mechanical tests. Specically, the crystalline peaks for the
s3 phase (the reinforcing dendrite phase) increase notably
with time-at-temperature for these composites. Furthermore, a third phase, with peaks matching those of the s5
phase (Zr38Cu36Al26), was found in some of the deformed
specimens with f = 20%.
Fig. 9 shows SEM images of two composite specimens,
f = 15% in Fig. 9a and f = 20% in Fig. 9b, after the longest
thermal exposures used in this work. These images can be
compared with Fig. 1c and d to appreciate the evolution

X.L. Fu et al. / Acta Materialia 55 (2007) 30593071

3065

Fig. 8. X-ray diraction patterns of as-cast as well as deformed samples with various initial reinforcement volume fractions, f = 0%, 7%, 15% and 20% in
(ad), respectively. Specimens that spent the most time-at-temperature were selected for this analysis. While the more dilute composites show no
discernible changes after deformation, the concentrated alloys in (c) and (d) show evidence of second-phase precipitation upon deformation at elevated
temperature.

of the structure during the high-temperature mechanical


test: the second-phase dendrites generally grow larger and
their volume fraction increases accordingly. Applying
image analysis techniques to these micrographs, we nd
that the volume percentage of the reinforcing phases
increased slightly to about 17% and 28% from f = 15%
and 20%, respectively. Although volume fraction measurements such as these are subject to some uncertainty, the
increased second-phase content suggested here is in line

with the evolution of the XRD peaks in Fig. 8, and attests


to a subtle microstructural evolution during the high-temperature mechanical testing.
Another possible contribution to the changes in the
XRD patterns of Fig. 8 is nanocrystallization within the
amorphous matrix. Such nanocrystallization has been
reported by prior authors during homogeneous deformation [23,34,36,4346] especially at temperatures around
the glass transition temperature (Tg) and higher stresses

Fig. 9. SEM micrographs of (a) f = 15% and (b) f = 20%, after undergoing uniaxial compression testing at T = 693 K at (a) 2 105 s1 (218 min at
temperature) and (b) 1 105 s1 (228 min at temperature). These micrographs can be compared with those in Fig. 1c and d, respectively, to perceive the
increase in second-phase reinforcement that apparently precipitates in situ during the high-temperature mechanical test.

3066

X.L. Fu et al. / Acta Materialia 55 (2007) 30593071

where non-Newtonian ow occurs. The f = 0% and 7%


specimens did not exhibit any measurable change in their
diraction patterns after deformation, and the more concentrated composites exhibited crystallization that was
nominally detected at the microstructural scale by SEM.
Therefore, we have no particular reason to believe that
there is a signicant amount of nanocrystallization in any
of the present deformed materials, although we cannot rule
out this possibility.
Although the above observations suggest that additional
precipitation of the second phase occurs upon reheating the
cast specimens to near Tg, these microstructural changes
are subtle, and apparently do not markedly impact the
structure or composition of the glass matrix. This is evidenced by the additional DSC scans we conducted on specimens previously deformed at high temperatures, which all
exhibited characteristic temperatures (Tg and Tx) within a
few degrees of the values obtained on the as-cast samples
of the same composition. This is depicted for T onset
in
g
Fig. 3b, where it is clear that, to within the uncertainty
on the measurement, the high-temperature deformation
and the accompanying microstructural evolution did not
aect the glass transition temperature.
4. Discussion
The trends displayed in Figs. 6 and 7 represent the rst
systematic exploration of mechanical properties in a single
class of in situ BMGMCs, as a function of temperature, rate
and reinforcement volume fraction. All of the tested materials exhibit similar responses with respect to temperature
and rate dependence, but of more interest presently are
the clear and systematic trends as the second-phase loading
increases. These trends are especially clear in the homogeneous regime, as captured in Figs. 6b and 7b. In the sections
that follow, we will explore the strengthening mechanisms
in these BMGMCs in the homogeneous regime, and quantitatively explain the trends in Figs. 6b and 7b. As a starting
point for this analysis, we begin by considering the response
of the unreinforced metallic glass composition.
4.1. Homogeneous deformation of the unreinforced metallic
glass
For the purpose of explaining trends in the deformation
of the base metallic glass used in this study (Zr49Cu45Al6,
f = 0%), we employ the phenomenology of Argons shear
transformation zone model [27]. In the steady-state homogeneous condition, this model provides a constitutive ow
law written in terms of uniaxial stress r and strain rate e_ as
[47]


r c0 V
e_ e_ 0 sinh p
1
3 kT
where k is Boltzmanns constant and T is temperature, and
e_ 0 captures the temperature dependence of ow (assuming
a strong glass former) via the activation energy Q



Q
e_ 0 A  exp 
kT

with A a pre-exponential constant in units of reciprocal


time. The parameter c0  0.125 characterizes the local
shear strain of the atomic-scale rearrangements (the socalled shear transformation zones) that accommodate
strain, and V is the characteristic volume associated with
such rearrangements. The data for the f = 0% alloy in Figs.
6b and 7b should be describable using Eq. (1), as they represent steady-state ow of a homogeneous metallic glass.
We will consider these two data sets in turn.
We rst examine the rate dependence of steady-state
ow, by tting Eq. (1) to the experimental data of
Fig. 6b (again focusing only upon the data at f = 0%, for
which an unmodied form of Eq. (1) is expected to apply),
using the method of least squares. Since all of these data
were acquired at the same temperature (T = 693 K), it is
not necessary (or fruitful) to consider the temperaturedependent details captured by Eq. (2), so we use instead
two free parameters in this t: the shear transformation
zone volume V and the amplitude e_ 0 . As shown by the solid
line in Fig. 6b, the data can be described quite well with this
model using e_ 0 2:46  104 s1 and V = 1.7 1027 m3.
While the former value is of no particular consequence in
the present discussion, the latter suggests a characteristic
size for shear transformation zones that is large enough
to contain about 70 atoms (assuming that the composition
of the shear transformation zone matches the global alloy
composition, with 60% packing density). This volume is
of about the same order as that suggested by Argon in
his original work on shear transformation zones in Pd80Si20
(V  8 1028 m3 [27]) and also comparable to more recent
estimates for the Zr-based metallic glass Vitreloy-1
(Zr41Ti14Cu12.5Ni10Be22.5, for which V  8 1028 m3
[47]), and for Pd- and Mg-based glasses (V  5 1028 m3
[25]). However, it is potentially signicant that the present
characteristic volume is about twice that estimated in the
prior works; we will return to this issue later.
Turning our attention now to the temperature eect on
steady-state ow as shown in Fig. 7b, we can again use
Eq. (1) and t the data, noting this time, however, that capturing the temperature dependence via Eq. (2) is in fact the
goal of the tting exercise and all of the data in Fig. 7b were
acquired at the same rate. Accordingly, we x the value of
V = 1.7 1027 m3 found above and treat this term as a
constant in the following. By least squares tting of the
experimental data points for the f = 0% alloy, Q is found
to be 660 kJ mol1. Changes in the assumed value of V
lead to variation in the resulting tted value of Q, but a similar magnitude is obtained for any reasonable variation in V.
The value of Q extracted from the experimental data
above can be expressed as a fraction of the thermal energy
at the glass transition, Q  115 kTg. This energy is at the
high end as compared with typical values for viscous ow
of metallic glasses, which lie in the range 20120 kTg
[25,41,47]. For comparison, the activation energy for the

X.L. Fu et al. / Acta Materialia 55 (2007) 30593071

3067

dln r e_ dr

dln e_ r d_e

Zr-based glass Vitreloy-1 has been found to be 445 kJ


mol1, or 85 kTg [47]; the present value is roughly 50%
greater than this.
As the above analysis shows, the present Zr49Cu45Al6
glass exhibits a relatively high activation volume (c0 V)
and a somewhat elevated activation energy as compared
with the similar Zr-based glass Vitreloy-1. The fact that
both of these deviations have the same sign is self-consistent; this may be seen more clearly by considering Argons
model for shear transformation zone activation [27]. This
model considers the elastic energy associated with distorting a local volume element of a continuum in a shear mode,
and gives the activation energy in terms of elastic constants


7  5m
21 m 2
c
Q

b c  l  c20  V
3
301  m 91  m
2c0
Here l and m are the shear modulus and Poissons ratio of
the glass, respectively, and cc  0.03 is the ideal shear strain
of the volume element. The parameter b  1 is the ratio of
volume dilatation to shear distortion associated with the
operation of the shear transformation zone.
Limiting our attention to a comparison of the present
glass and Vitreloy-1, we note that Poissons ratio is virtually identical for the two materials [48]. Together with the
assumption that the characteristic strain c0 is of the same
order in both glasses, this renders the bracketed term in
Eq. (3) essentially irrelevant, and we see that the activation
energy scales simply as Q  l V. Unrelaxed shear moduli
in the regime of homogeneous ow are not available for the
present glass, but the considerations in Ref. [48] suggest
that the ratio of moduli (Zr49Cu45Al6 vs. Zr41Ti14Cu12.5Ni10Be22.5) is about 0.8. Introducing the estimated characteristic volumes V  1.7 and 0.8 nm3 for Zr49Cu45Al6
and Vitreloy-1, respectively, this analysis suggests that
the activation energy for the present glass should actually
be about 70% higher than that in Vitreloy-1. Indeed, our
experimental measurements suggest a dierence of about
50%, which is quite reasonable agreement. Thus, the larger
activation energy in our glass is commensurate with the larger characteristic volume of the shear transformation zones
that carry the deformation.
The above calculation shows that the dierences in ow
behavior between Zr49Cu45Al6 and Vitreloy-1 stem from
dierent characteristic ow defects in the two glasses, i.e.
shear transformation zones of measurably dierent sizes
and energies. Although both glasses are based on Zr, they
have substantially dierent compositions, so this dierence
is not necessarily surprising or unexpected. Further speculation as to the structural origins of this dierence are
beyond our present scope, but clearly present opportunities
for future research.
One nal point pertaining to the homogeneous deformation behavior of amorphous Zr49Cu45Al6 that will be of
importance in understanding the rheology of our in situ
composites is the strain-rate sensitivity. This can be discussed quantitatively through consideration of the strain
rate exponent m

which is equal to unity for Newtonian ow (i.e. strong linear rate dependence) and decreases as the deformation becomes less strain-rate-sensitive. When Eq. (4) is evaluated
using the constitutive ow law of a metallic glass in steady
state (Eq. (1)), a transition from Newtonian to non-Newtonian deformation is predicted as stress increases. This
behavior is illustrated in Fig. 6b at T = 693 K, where the
slope of the trendlines directly gives the strain-rate sensitivity m. At low stresses, Fig. 6b shows a series of curves that
converge to a slope of m  1, with steadily decreasing rate
sensitivity at higher stresses.
Outside of the Newtonian regime (which is promoted at
low stresses and higher temperatures), the strain-rate sensitivity varies rapidly with both applied stress and temperature. For example, the data in Fig. 7b show the steadystate ow strength of our glass as a function of temperature
and, based on Eqs. (1) and (4), we nd that none of these
measurements lie in the Newtonian regime. In fact, as the
temperature decreases and the strength of the glass rises
in Fig. 7b, we nd that the strain-rate sensitivity is below
m = 0.1. As we will see shortly, this result has some bearing
on the mechanics of composite deformation discussed in
the following section.
4.2. Homogeneous deformation of in situ BMG composites
The above section illustrates that the steady-state homogeneous deformation of the base amorphous alloy (unreinforced) is well described by Argons model, including both
Newtonian and non-Newtonian response, over a range of
rates and temperatures. This is important for understanding the case of the reinforced BMGMCs, because we expect
that the deformation is likely carried solely by homogeneous deformation of the amorphous matrix phase, which
is near Tg and soft compared with the intermetallic reinforcements. Accordingly, the same basic ow mechanisms
that operate in the unreinforced glass are expected to occur
in the matrix of the composites.
4.2.1. Strengthening mechanisms in the Newtonian regime
A rst point of comparison between the composites and
the unreinforced BMG is provided by considering the rate
dependence of homogeneous ow, as shown in Fig. 6b.
Here we see a clear strengthening eect that occurs upon
addition of the reinforcement, although the curves all exhibit the same general shape, with a transition from nearly
Newtonian to non-Newtonian behavior as the stress is
increased. If the reinforcements are assumed to be rigid
(which is reasonable for the relatively low homologous
temperatures they experience), then the rate dependence
seen in Fig. 6b must be ascribed solely to the creep rheology of the surrounding glassy matrix. Because the composites are all based on the same essential glass composition,
one would also expect in turn that the deformation

3068

X.L. Fu et al. / Acta Materialia 55 (2007) 30593071

mechanism (i.e. the characteristic shear transformation


zone) should be similar in all of these materials. This is veried by considering the value of the characteristic volume,
V, and activation energy, Q, for each of these composites.
In Eq. (1), variations in V amount to changes in the curvature of data such as those in Fig. 6b, while changes in the
amplitude e_ 0 lead to shifts in the position of the curve. The
data series in Fig. 6b are shifted relative to one another
(reective of a change in e_ 0 but all exhibit quite similar
curvature (reective of a relatively constant value of V).
Indeed, by least-squares tting each of these data sets separately, we nd that the value of V is reasonably constant
with composition. This result is also supported by the
results in Fig. 7b, which show that the temperature dependence of deformation is roughly the same in all of our
experimental materials. In fact, more detailed tting of
these data show that the activation energy, Q, is roughly
a constant across composition, which is in line with a constant value of V via Eq. (3).
In light of the above discussion, it is reasonable to say
that, with only a single adjustable parameter _e0 , we can
convincingly t the ow data of the composites by assuming that the same basic deformation mechanism occurs in
the glassy matrix of each. This result is illustrated by the
dashed lines in Fig. 6b, which plot the tted form of Eq.
(1) using the same characteristic volume V = 1.7 nm3
obtained earlier for the unreinforced glass. The resulting
tted values of e_ 0 are 8.0 105, 2.0 105 and 2.6
106 s1, respectively, for f = 7%, 15% and 20%. The
decrease in e_ 0 with f directly quanties the strengthening
eect that arises with reinforcement loading, and we may
now consider the strengthening mechanisms that lead to
this decrease. In Fig. 10, the parameter e_ 0 is normalized
to that obtained in the unreinforced alloy _ef0 0 and collected as a function of the reinforcement of volume
fraction.
For the purposes of analyzing these data, it is useful to
make the following observation: in the limit of low applied
stresses, the amplitude e_ 0 corresponds to the reciprocal viscosity, i.e. Eq. (1) reduces as
e_ / e_ 0  r

This extrapolation is signicant, because in the Newtonian


limit there is a closed-form model by Lee and Mear [49]
that predicts composite strengthening. This model is a continuum treatment that describes load sharing between matrix and reinforcement, and can predict the global response
of the composite. Assuming equiaxed reinforcements, their
model gives
d
e_ c0 e_ m
0  1  f

where the superscripts c and m denote composite and matrix properties, respectively, and the exponent d = 5/2 for a
Newtonian ow condition.
The predictions of Eq. (6) are plotted in Fig. 10 as a
dashed line. The generally poor agreement between this
model and the experimental data suggest that typical com-

Fig. 10. Characteristic Newtonian ow rates at T = 693 K for the in situ


composites specimens, normalized by that of the unreinforced glass. The
symbols are the same as Fig. 6b, and the points were extracted from the
data of Fig. 6b through best tting with Eq. (1). The downward trend of
the data with increasing f is reective of strengthening, as modeled by the
load-transfer model of Eq. (6) (dashed line), as well as the modied form
of Eq. (8), which further incorporates the eects of matrix strengthening
via a shift the matrix glass transition temperature, Tg (solid line). The
dotted line illustrates the eect of matrix strengthening without any
consideration of load-transfer (Eq. (8) with d = 0). For the two more
concentrated composites, some amount of precipitation occurred during
deformation (cf. Figs. 8 and 9), leading to subtle increases in f as noted by
the arrows in this plot. Some of the data in this gure appeared previously
in Ref. [41].

posite strengthening through load transfer does not


account for the high strength of these BMGMCs. The reason for this disagreement lies in an implicit assumption that
we have made up to this point, namely, that the matrix
properties of the composite specimens are adequately represented by our measurements on the unreinforced glass
_ f0 0 ). In fact, as noted earlier, these
(i.e. that e_ m
0 is equal to e
BMGMCs each have slightly dierent matrix structures
and/or compositions, as reected in the shifted values of
Tg shown in Fig. 3a and b. With increasing aluminum content, the composites have higher values of Tg, and thus the
mechanical tests were conducted at somewhat lower
homologous temperatures (T/Tg). This is tantamount to a
direct chemical or structural strengthening of the glassy
matrix in these composites. Consequently, e_ m
0 is decidedly
not equal to e_ f0 0 , because of the inherent temperature
dependence carried in Eq. (2).
It is perhaps a fortunate coincidence that Tg shifts in a
systematic, nearly linear fashion with respect to f, and we
can write the empirical equation
f 0
Tm
vf
g  Tg

where the matrix of a composite is now dierentiated from


the unreinforced glass by the superscripts, and the coecient v  65 K derives from the data in Fig. 3b. Combining
this expression with Eq. (2), we can write a modied form
of Eq. (6)


Q0  v  f
d
e_ c0 e_ f0 0  1  f  exp 
8
T

X.L. Fu et al. / Acta Materialia 55 (2007) 30593071

where Q 0 = 115 is the activation energy expressed in units


of kTg. Eq. (8) now includes two composition-dependent
terms: the rst is the composite load-sharing eect derived
by Lee and Mear (Eq. (6)), and the second is a term
accounting for the shift in matrix composition upon precipitation of the in situ reinforcements.
Eq. (8) matches well to the experimental data shown in
Fig. 10, and by comparing the solid line to the dashed one
(Eq. (8) vs. Eq. (6), respectively) it is clear that the dominant contribution to strengthening is the shift in matrix
composition from one composite to the next. This is also
observed by setting d = 0 in Eq. (8) and plotting the eect
of matrix strengthening directly as the dotted line in
Fig. 10. While the matrix composition shift in the composites does not markedly change the deformation mechanism
(i.e. Q or V), the subtle evolution of the glass transition
temperature leads to dramatic strengthening. The reason
for this is the very high temperature dependence of strength
(cf. Fig. 7) in the high-temperature regime; small variations
in homologous temperature can lead to dramatic changes
in ow behavior.
A nal interesting point about Fig. 10 pertains to the
evolution of the composite structures during deformation.
As observed earlier, the f = 15% and 20% composites
experienced additional phase separation when exposed to
high-temperature testing conditions, leading to subtle but
tangible increases in the eective volume fraction of reinforcing phases (see Figs. 8 and 9). The two data points
labeled with arrows in Fig. 10 illustrate the eect of this
evolution on the experimental data; for the same set of
mechanical test data, the volume fraction of reinforcement
can be regarded as evolving towards the right of the plot.
Interestingly, this evolution can explain the minor discrepancies between the predictions of Eq. (8) and the data, providing surprisingly good agreement between them.
4.2.2. Strengthening in the non-Newtonian regime
In our earlier analysis, we discussed mechanisms of
strengthening including most notably (i) composite load
sharing, and (ii) matrix strengthening by virtue of composition shift. In analyzing the data of Fig. 7b, we can more
cleanly separate these two eects, by directly correcting
the data for the shift in Tg. We do this by tting each
data set with the dominant temperature dependence of
Eq. (2), assigning Q  115 kTg to directly incorporate
the measured values of Tg for each composition. The
composite strengthening eect is then revealed by examining the values of the tted pre-exponential constants,
A, for each data set. In Fig. 11, we illustrate the eect
of composite reinforcement loading on A (which is normalized to the value for the unreinforced composite,
Af=0, in analogy to Fig. 10). We emphasize that by taking Q 0 = 115 as a constant, as discussed above, the eects
of matrix strengthening (i.e. shifts in Tg) are directly
accounted for in this analysis, and the results in Fig. 11
show strengthening due only to composite load-sharing
eects.

3069

Fig. 11. Composite strengthening eects derived from the temperaturedependent data in Fig. 7b, obtained by tting to the pre-exponential
constants A (in Eq. (2)) for each specimen. These data are normalized to
the response of the unreinforced glass (f = 0%), and directly account for
the dierent glass transition temperatures of the various composites.
Accordingly, the strengthening behavior with f seen here is due solely to
load-transfer eects, which is modeled by the composite model of Eq. (9).
Accounting for the low rate-sensitivity of these data (low m values near
0.1), the model adequately matches the data. As in Fig. 10, the structural
evolution (shift in f) during testing is denoted by the arrows.

As Fig. 11 shows, even when the eect of matrix


strengthening is directly removed from the experimental
data, the composite strength increases with the volume
fraction of reinforcements. We may write a modied form
of Eq. (6) as
Ac Af 0  1  f

where the expression is written in terms of A instead of e_ 0 to


explicitly remove thermal eects; A is temperature-independent. As described earlier, for a state of Newtonian ow
(i.e. m = 1), the constant d = 5/2. However, the data in
Fig. 7b were not acquired under Newtonian ow conditions, as described previously in Section 4.1, where m values ranging as low as 0.1 were calculated for these data.
Therefore, it is perhaps not surprising that Eq. (9) underpredicts the level of composite load sharing we see in the
experimental data when the Newtonian condition is assumed (m = 1, d = 5/2), as shown by the dashed line in
Fig. 11.
Unfortunately, for a matrix obeying a nonlinear constitutive law such as Eq. (1), we know of no closed-form analytical solution to the composite deformation problem.
However, Lee and Mear [49] did present a numerical analysis of composite creep that considered nonlinear matrix
behavior, and tabulated values of d for various exponents,
m. In the present case, where m ranges from 0.08 to 0.33,
the appropriate values of d lie in the range 36 [49]. This
signicantly increases the amount of composite strengthening predicted by Eq. (9), as shown by the series of curves in
Fig. 11. The solid line in Fig. 11 shows that, given suciently nonlinear matrix behavior (m  0.1), the strength
of the experimental data is reasonably reproduced by Eq.
(9). And, as seen earlier in Fig. 10, the two data points

3070

X.L. Fu et al. / Acta Materialia 55 (2007) 30593071

labeled with arrows show the composite structure evolution


we have observed in these tests; this eect oers a further
mechanism that can bring the experimental data in line
with the predictions of Eq. (9).
4.2.3. Implications
The above sections show that (i) load sharing and (ii)
matrix hardening are the two sources of strengthening in
the present BMGMCs. Depending upon the test conditions, the relative importance of these two eects may vary.
For example, our data acquired at T = 693 K (near Tg of
the unreinforced alloy) in the Newtonian regime (Fig. 10)
illustrate a strong apparent strengthening due to the shift
in matrix composition and a relatively inconsequential
composite load sharing eect. On the other hand, the data
in Fig. 11 are already corrected for the matrix hardening
eect, and the strengthening observed here is due entirely
to the load-sharing eect. Our analysis shows that the
strain-rate sensitivity is likely the dominant factor controlling this selection: all other eects being equal, high rate
sensitivity tends to increase the importance of composite
load sharing.
The above considerations lead to an important practical
conclusion for BMGMCs: the transient strength of a BMG
in the non-Newtonian regime will be impacted more certainly by the incorporation of reinforcements than will
the steady-state ow behavior. Therefore, if the matrix
composition can be properly maintained upon precipitation of the reinforcement phase (i.e. if matrix hardening
can be ruled out), then a BMGMC will maintain the formability of the parent glass with virtually no strengthening in
the Newtonian regime, while deriving the potential benets
of higher strength, improved ductility and more predictable
properties at lower temperatures.
In principle, the strengthening eects that we have discussed in this work should be common to all in situ
BMGMCs, although the relative importance of load sharing, matrix hardening and second-phase precipitation during deformation may certainly be expected to vary from
one family of glasses to another. For example, it is easy
to conceive a BMGMC in which the matrix softens upon
precipitation of the second phase rather than strengthening. This would occur if the homologous temperature
(T/Tg) increases as the second phase precipitates. The
developments in this work also pertain specically to the
case in which the reinforcements are rigid with respect to
a deforming matrix; in some materials where, for example,
pure metals precipitate from the glass, it is certainly plausible that both phases might deform together, rendering the
load-sharing problem considerably more complex.
A nal issue which limits the scope of the present work
is the fact that the present BMGMCs are all reasonably
dilute. Even at the highest volume fractions we have investigated, the volume percentage of the second phase is below
the typical percolation thresholds of three-dimensional
composites, and therefore the average global behavior of
the composite is dominated by the properties of the matrix

phase. At higher volume fractions in the vicinity of the


percolation threshold, the properties of the composite will
represent a more complex average of the phase properties,
and at suciently high f above the threshold, the properties
will be dominated by the interconnected second phase.
In room-temperature mechanical tests on La-based
BMGMCs, Lee et al. [13] have shown the critical importance of phase connectivity: when f reached a critical value
near 40%, rapid changes in ductility and impact toughness
manifested. Some of our future research will consider the
deformation of more concentrated BMGMCs in the hightemperature homogeneous regime.
5. Conclusions
The deformation behavior of a Zr-based BMG and
three derivative in situ composites has been systematically
studied over a wide range of strain rates and temperatures.
Particular emphasis has been placed upon the transition to
homogeneous deformation at high temperatures, and rheology near the glass transition. The eect of second-phase
reinforcements on ow has been specically investigated
in both Newtonian and non-Newtonian ow regimes.
The following main conclusions have been drawn from this
investigation:
 The same basic modes of deformation were observed in
both the Zr-based BMG and the composites studied in
this work (with volume fractions of 7%, 15% and 20%
reinforcing intermetallic phase). Inhomogeneous ow
and apparent brittle fracture were observed at low temperatures, with a transition to homogeneous ow at high
temperatures. At combinations of stress and temperature close to the transition region, stress overshoot was
observed before reaching a steady-state ow condition.
In the homogeneous regime, deformation becomes very
sensitive to both temperature and strain rate.
 In the homogeneous regime, as the volume percentage of
the second phase increases, the ow stress increases (i.e.
the steady-state strain rate decreases). The mechanism
for this strengthening has been explored by combining
Argons shear transformation zone model to describe
the homogeneous ow of the matrix with a composite
reinforcement model to capture load sharing between
the creeping matrix and rigid particles. This analysis
reveals that the strengthening eect arises from two separate contributions: load transfer from the amorphous
matrix to the reinforcements and direct hardening of
the glassy matrix due to changes in the glass composition and structure upon precipitation of reinforcements.
This conclusion is supported quantitatively for both
Newtonian and non-Newtonian ow data.
 In some of the composites with higher volume fractions
of in situ reinforcement, additional second-phase precipitation was observed during deformation. This led to an
increase in the eective volume fraction of reinforcement, and additional strengthening.

X.L. Fu et al. / Acta Materialia 55 (2007) 30593071

Acknowledgements
This work was supported by the SingaporeMIT Alliance, as well as the US Army Research Oce under contract DAAD19-03-1-0235; the views expressed in this
work are not endorsed by the sponsors. The authors gratefully recognize the experimental involvement of Dr. M.L.
Lee and Mr. D. Wang (National University of Singapore),
as well as Mr. A.J. Detor (MIT).
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]

Johnson WL. MRS Bull 1999;24:42.


Telford M. Mater Today 2004;7:36.
Huang JC, Chuang TH. Mater Chem Phys 1999;57:195.
Inoue A. Acta Mater 2000;48:279.
Bae DH, Lee MH, Kim DH, Sordelet DJ. Appl Phys Lett
2003;83:2312.
Conner RD, Choi-Yim H, Johnson WL. J Mater Res 1999;14:3292.
He G, Loser W, Eckert J, Schultz L. J Mater Res 2002;17:3015.
Fan C, Ott RT, Hufnagel TC. Appl Phys Lett 2002;81:1020.
Choi-Yim H, Conner RD, Szuecs F, Johnson WL. Acta Mater
2002;50:2737.
Szuecs F, Kim CP, Johnson WL. Acta Mater 2001;49:1507.
Hirano T, Kato H, Matsuo A, Kawamura Y, Inoue A. Mater Trans
JIM 2000;41:1454.
Bian Z, Ahmad J, Zhang W, Inoue A. Mater Trans 2004;45:2346.
Lee ML, Li Y, Schuh CA. Acta Mater 2004;52:4121.
Das J, Loser W, Kuhn U, Eckert J, Roy SK, Schultz L. Appl Phys
Lett 2003;82:4690.
Hays CC, Kim CP, Johnson WL. Phys Rev Lett 2000;84:2901.
Wang G, Shen J, Sun JF, Stachurski ZH, Zhou BD. Acta Metall Sin
2005;41:291.
Kawamura Y, Nakamura T, Inoue A. Scr Mater 1998;39:301.
Heggen M, Spaepen F, Feuerbacher M. Mater Sci Eng A Struct
Mater Prop Microstruct Process 2004;375377:1186.
Spaepen F. Acta Metall 1977;25:407.
Argon AS, Shi LT. Acta Metall 1983;31:499.
Perez J. Acta Metall 1984;32:2163.

3071

[22] Schuh CA, Argon AS, Nieh TG, Wadsworth J. Philos Mag
2003;83:2585.
[23] Nieh TG, Wadsworth J, Liu CT, Ohkubo T, Hirotsu Y. Acta Mater
2001;49:2887.
[24] Chu JP, Chiang CL, Mahalingam T, Nieh TG. Scr Mater
2003;49:435.
[25] Schuh CA, Lund AC, Nieh TG. Acta Mater 2004;52:5879.
[26] Polk DE, Turnbull D. Acta Metall 1972;20:493.
[27] Argon AS. Acta Metall 1979;27:47.
[28] Argon AS, Kuo HY. Mater Sci Eng 1979;39:109.
[29] Megusar J, Argon AS, Grant NJ. Mater Sci Eng 1979;38:63.
[30] Patterson J, Jones DRH. Scr Metall 1979;13:949.
[31] Gibeling JC, Nix WD. Scr Metall 1978;12:926.
[32] Nieh TG, Wadsworth J. Scr Mater 2006;54:387.
[33] Spaepen F. Scr Mater 2006;54:363.
[34] Wang G, Shen J, Sun JF, Huang YJ, Zou J, Lu ZP, et al. J NonCryst Solids 2005;351:209.
[35] Bletry M, Guyot P, Brechet Y, Blandin JJ, Soubeyroux JL.
Intermetallics 2004;12:1051.
[36] Nieh TG, Schuh C, Wadsworth J, Li Y. Intermetallics 2002;10:1177.
[37] Chu JP, Chiang CL, Nieh TG, Kawamura Y. Intermetallics
2002;10:1191.
[38] Lu J, Ravichandran G, Johnson WL. Acta Mater 2003;51: 3429.
[39] Bae DH, Lee MH, Yi S, Kim DH, Sordelet DJ. J Non-Cryst Solids
2004;337:15.
[40] Bae DH, Park JM, Na JH, Kim DH, Kim YC, Lee JK. J Mater Res
2004;19:937.
[41] Fu XL, Li Y, Schuh CA. Appl Phys Lett 2005:87.
[42] Wang D, Tan H, Li Y. Acta Mater 2005;53:2969.
[43] Wang G, Fang SS, Xiao XS, Hua Q, Gu HZ, Dong YD. Mater Sci
Eng A Struct Mater Prop Microstruct Process 2004;373:217.
[44] Shen J, Wang G, Sun JF, Stachurski ZH, Yan C, Ye L, et al.
Intermetallics 2005;13:79.
[45] Bae DH, Lim HK, Kim SH, Kim DH, Kim WT. Acta Mater
2002;50:1749.
[46] Chiang CL, Chu JP, Lo CT, Nieh TG, Wang ZX, Wang WH.
Intermetallics 2004;12:1057.
[47] Schuh CA, Hufnagel TC, Ramamurty U. Acta Mater, submitted for
publication.
[48] Wang WH. J Appl Phys 2006;99:093506.
[49] Lee BJ, Mear ME. J Mech Phys Solids 1991;39:627.

S-ar putea să vă placă și