Sei sulla pagina 1di 11

Acta Materialia 52 (2004) 41214131

www.actamat-journals.com

Eect of a controlled volume fraction of dendritic phases on


tensile and compressive ductility in La-based metallic glass
matrix composites
M.L. Lee a, Y. Li
a

a,*

, C.A. Schuh

Department of Materials Science, National University of Singapore, Lower Kent Ridge Road, Singapore 119260, Singapore
b
Department of Materials Science and Engineering, MIT, Cambridge, MA, USA
Received 23 February 2004; received in revised form 13 May 2004; accepted 14 May 2004
Available online 17 June 2004

Abstract
A systematic study has been carried out to identify the eect of crystalline second phase reinforcements on the mechanical
properties of amorphous alloys based on the composition La86y Al14 (Cu, Ni)y (y 124). By varying the composition in a controlled manner, composites with a tailorable fraction of in situ dendritic reinforcements were produced. Tension, compression and
impact tests showed considerably improved mechanical properties with a large amount of plasticity when the reinforcement volume
fraction exceeded a critical value near 40%, and the tension/compression asymmetry associated with glass plasticity is reduced as the
volume fraction of crystalline reinforcements rises.
 2004 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Keywords: Mechanical properties; Bulk amorphous materials; Composites

1. Introduction
In the past few years, a new class of composite
materials has emerged, combining an amorphous metallic matrix with metallic or ceramic reinforcements of
various morphologies [111]. The rst such materials
were formed through an ex situ process by which solid
crystalline phases were added to the molten matrix [1
3]. Later, several groups developed in situ composites
in which the reinforcement phase nucleates from a solid
solution, most often during the process of cooling from
the melt [48,10,11]. Beginning with the work of Hays
et al. [6], these in situ composites have now been developed in several bulk metallic glass (BMG) forming
systems. In these composites, the BMG matrix provides
extreme strength, while the presence of reinforcing
phases can apparently suppress catastrophic failure due
*

Corresponding author. Tel.: +65-687-433-48; fax: +65-677-636-04.


E-mail address: masliy@nus.edu.sg (Y. Li).

to shear localization, and leads to legitimate plastic


ow [2,46,10]. The resulting unique combination of
ceramic-like strength with metal-like ductility is a
strong technological motivation for further work in this
area.
Although prior works have demonstrated the benet
of coarse crystalline reinforcements in metallic glasses,
there has yet to be a systematic study of the role of reinforcement volume fraction on various properties, especially for the new class of in situ composites.
Therefore, it remains unclear which properties can be
reasonably described using, e.g., rule-of-mixtures type
approaches, and which properties require new physical
models. The purpose of this paper is to present the rst
systematic study of volume-fraction eects on the mechanical properties of model in situ BMG composites.
Using specimens with a wide range of reinforcement
loadings, we examine yield and fracture stresses, Charpy
impact response, fracture surfaces, as well as evidence
for plastic asymmetry.

1359-6454/$30.00  2004 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.actamat.2004.05.025

4122

M.L. Lee et al. / Acta Materialia 52 (2004) 41214131

2. Experimental procedure
Monolithic La62 Cu12 Ni12 Al14 amorphous alloy and
La86y Al14 (Cu, Ni)y (y 120) alloy composites were
prepared by arc-melting a mixture of La (99.9%), Al
(99.9%), Ni (99.98%) and Cu (99.999%) in an argon
atmosphere. Since pure a-La is known to precipitate
from this system when the La content is increased [11], a
systematic variation of composition (via a change in y)
can produce BMG composites with a chosen volume
fraction of reinforcing a phase. The BMG alloy and its
composites were prepared by remelting the master ingots with various selected values of y at a temperature
of 973 K in an argon atmosphere and casting into a
copper mould with a 5 mm diameter cavity. Cross-sections of the rods were examined by X-ray diraction
(XRD) and the glass transition and crystallization
temperatures were identied with a dierential scanning
calorimeter (DSC) at a heating rate of 20 K/min. Standard image analysis techniques were used on scanning
electron micrographs to determine the volume fraction
of dendrites in the composites. For comparison with
these glassy alloys and composites, a specimen of pure
crystalline La was also produced and characterized by
the same methods.
Non-standard Charpy specimens, 60 mm in height
and 5 mm in diameter, were prepared and notched
(notch depth of 0.5 mm) with a low speed diamond saw.
The Charpy impact tests were carried out with an instrumented Charpy test machine. The drop hammer of
the impact machine was calibrated to an impact energy
of 0.657 J at a velocity of 1.210 m/s, and the resulting
impact fracture energy was recorded for each of the
specimens.
An Instron 5500R load frame was used to test three
specimens of each composition under uniaxial loading at
room temperature. The compression test specimens were
10 mm in length and 5 mm in diameter, providing a
nominal length-to-diameter ratio of 2:1 as recommended
by ASTM E9-89a for testing high strength material. The
compression specimens were ground parallel using SiC
grinding papers followed by polishing with alumina
powder, to an accuracy of less than 10 lm. The compression samples were sandwiched between two WC
platens in a loading xture designed to guarantee axial
loading. The ends of the compression samples were lubricated to prevent barreling of the samples. Strain gages (TML) were glued on the surface of the specimens
gage section to obtain one-dimensional surface strains.
The tensile test specimens were machined into dog-bone
geometry with dimensions proportional to ASTM
standard EM8-01. The tensile specimens had a nominal
gage diameter of 2.5 mm and gage length of 15 mm. The
gage section of each specimen was also polished using
SiC paper and alumina powder (3 lm). A video extensometer was used to measure the strains during ten-

sile testing. All compression and tensile tests were


conducted using a constant strain rate of 104 s1 .

3. Results
3.1. Materials characterization
3.1.1. Microstructure
The scanning electron micrographs in Fig. 1 show
characteristic microstructures obtained in this study.
The amorphous alloy exhibits a featureless microstructure, while the BMG composite structures consist of a
uniformly distributed crystalline dendritic phase and an
amorphous matrix. Based on quantitative image analysis, we nd that the volume fraction of crystalline phases
increases from 0% to 50% as the value of y decreases
from 24 to 1 along the La86y Al14 (Cu, Ni)y line (see
Table 1). Although appropriate changes in the processing conditions can eect dierent size dendrites, the
length of the primary dendrite axes have been measured
to be in the same range for all of these materials, generally 412 lm; the eect of reinforcement size on mechanical behavior will be the subject of future research.
3.1.2. XRD analysis
XRD patterns of the monolithic La-based amorphous metal and composites are shown in Fig. 2(a). The
pattern of the monolithic alloy exhibits the broad peak
that is characteristic of the amorphous structure. The
diraction patterns of the composites show intense
crystalline peaks corresponding to hcp a-La phases,
superimposed upon the broad diraction maxima of the
amorphous phase. The broad diraction peak becomes
less obvious with increasing amount of crystalline dendrites in the matrix. The pure La sample was polycrystalline and its XRD scan exhibits crystalline peaks
corresponding to the same hcp a-La phase as the reinforcing dendrites.
3.1.3. DSC results
Fig. 2(b) shows the DSC results of all the alloys
studied. The values of the glass transition (Tg ) and
crystallization (Tx ) temperatures are summarized in Table 1. The glass transition temperatures were similar for
all of the alloys with variation of not more than 15 C.
A stronger dierence between the DSC curves of the
alloys lies in their crystallization peaks; Tx increases
slightly with increasing volume fraction of dendrites.
The rst crystallization peak shown on the DSC scan of
the monolithic BMG is attributed to the crystallization
of a-La phases as has been reported by Tan et al. [12].
Here we conrm the same essential behavior in the DSC
scan of the composite-forming alloys with y 20, which
exhibits the same rst crystallization peak corresponding

M.L. Lee et al. / Acta Materialia 52 (2004) 41214131

4123

Fig. 1. Backscattering SEM images of polished and chemically etched cross-sections of monolithic amorphous alloy and composites. The second
phases appear dark, where the fully amorphous matrix phase appears bright. The compositions are La86y Al14 (Cu, Ni)y with (a) y 24, Vf  0, (b)
y 20, Vf  0:07, (c) y 16, Vf  0:37 and (d) y 12, Vf  0:50.

Table 1
Results of DSC analysis for La86y Al14 (Cu, Ni)y (y 124) at heating rate of 20 K/min
Alloys

Vf

Tg (K)

Tx (K)

Tm (K)

Tl (K)

DTx (K)

Trg

DHx (J/g)

La62 Al14 (Cu,Ni)24


La66 Al14 (Cu,Ni)20
La70 Al14 (Cu,Ni)16
La71 Al14 (Cu,Ni)15
La72 Al14 (Cu,Ni)14
La74 Al14 (Cu,Ni)12
La85 Al14 (Cu,Ni)1

0
0.07
0.37
0.37
0.41
0.50
0.53

413.63
409.23
421.43
422.56
426.72
426.67

445.46
435.23
479.60
478.25
479.27
474.13
567.93

671.79
667.92
667.07
666.24
667.67
666.17

732.24
667.92
753.76
768.37
784.73

31.83
26.00
58.17
55.69
52.55
47.46

0.565
0.613
0.559
0.549
0.543

46.17
39.76
26.96
32.91
26.41
17.67
4.810

to La precipitation from the glass phase. As the (Ni,Cu)


content decreases with y, more La is precipitated as
crystalline dendrites during casting, and is therefore not
available to precipitate from the amorphous phase
during the DSC scans. Accordingly, the alloys with
y < 20 at.% showed only the secondary crystallization
peaks. Although the details of the crystallization events
observed in DSC therefore dier somewhat with composition, the results in Fig. 2(b) are consistent with the
very dierent volume fractions of a-La observed in the
microstructures of Fig. 1.
3.2. Mechanical properties
Fig. 3(a) shows representative uniaxial tensile stress
strain curves for the composites and amorphous samples. All the test results are summarized in Table 2,

together with data for a pure a-La sample acquired for


comparison with the BMG composites. The monolithic
BMG sample failed catastrophically showing a linear
elastic behavior up to a fracture stress of 550 MPa.
With a low volume fraction of reinforcement (e.g. 7%),
the composites showed very similar behavior, with no
obvious sign of plastic ow. At higher values of Vf , the
stressstrain curves start to show some evidence of
plastic yielding before failure; beyond Vf  40%, extensive plastic ow and work hardening are apparent.
Similar compressive stressstrain behavior can also be
seen for the monolithic BMG and composite samples in
Fig. 3(b); higher reinforcement loading leads to increased compressive ductility but lower yield strength.
The tabulated Charpy impact test data in Table 2 also
revealed a signicant improvement in the toughness of
the materials at higher Vf .

4124

M.L. Lee et al. / Acta Materialia 52 (2004) 41214131


(002)
(101)

(110)

(103)

(112)

(004)

Pure La
y = 1, Vf ~ 0.53
y = 12, Vf ~ 0.50

Intensity

Vf = 0.07

Vf = 0

600

y = 14, Vf ~ 0.41
y = 15, Vf ~ 0.37
y = 16, Vf ~ 0.37

Vf = 0.41

500
Tensile Stress (MPa)

(100)

Vf = 0.50

400

300

200

y = 20, Vf ~ 0.07

100
y = 24, Vf ~ 0

(a)
20

40

(a)

60

600

y = 14, Vf ~ 0.41

Exothermic

y = 15, Vf ~ 0.37
y = 16, Vf ~ 0.37
y = 20, Vf ~ 0.07

Compressive Stress (MPa)

y = 12, Vf ~ 0.50

Fig. 2. (a) XRD and (b) DSC scans of monolithic amorphous and
composite samples of La86y Al14 (Cu, Ni)y (y 124).

Measurements on the pure a-La samples also showed


high ductility, with a yield stress of only about 60 MPa.
After extended work hardening, an ultimate strength of
100 MPa at a tensile plastic strain of about 9% was
measured. The compressive test runs were stopped at
total strain of more than 15% without reaching fracture,
as the test specimens were losing their cylindrical shape.

4. Discussion
In the present system, we have successfully produced
in situ BMG composites that dier primarily in their
reinforcement volume fractions, with nominally similar
microstructural length scales and glass transition temperatures of the matrix. Therefore, it is reasonable to
assume that the dierent mechanical behaviors we ob-

0.41

0.50

500

0.53

400

300

200

0
(b)

600

Vf = 0

100

y = 24, Vf ~ 0

500
Temperature (K)

0.07

y = 1, Vf ~ 0.53

400

80

2 Theta

(b)

3
4
Strain (%)

Strain (%)

Fig. 3. Comparison of representative (a) tensile and (b) compression


results between monolithic amorphous alloy and composite samples.

serve among these specimens are most dominantly affected by the dendrite volume fraction, and to examine
these behaviors systematically on this basis. In the following sections we discuss the strength, ductility, and
impact toughness of these materials as a function of Vf ,
and later discuss asymmetry in the yield behavior and
fracture mechanisms as well.
4.1. Critical volume fraction for ductile behavior
The relation between the compressive/tensile yield
and fracture strength with volume fraction of dendrites
is plotted in Fig. 4, using as endpoints the data for the
monolithic glass specimen (Vf 0) and the pure a-La
specimen (Vf 1). The trends in Fig. 4 suggest that
despite some scatter (especially in the compressive data)
the yield and fracture strength of the composites can be
reasonably described by a rule-of-mixtures or, in principle, by more complex composite mechanics models
based on load transfer.

M.L. Lee et al. / Acta Materialia 52 (2004) 41214131

4125

Table 2
Summary of tensile, compressive and charpy impact test data
Specimen

Vf

La62 Al14 (Cu,Ni)24


La66 Al14 (Cu,Ni)20
La70 Al14 (Cu,Ni)16
La71 Al14 (Cu,Ni)15
La72 Al14 (Cu,Ni)14
La74 Al14 (Cu,Ni)12
La85 Al14 (Cu,Ni)1
La

0
0.07
0.37
0.37
0.41
0.50
0.53
1.0

Tensile test

Compression test

Charpy impact test

r0:2%
(MPa)

rf
(MPa)

eplastic
(%)

r0:2%
(MPa)

rf
(MPa)

eplastic
(%)

Impact
toughness
(kJ/m2 )

AfShear

354
355
322
352

60

549
592
490
459
460
435

79

0
0
1.3
1.0
1.3
5.0

9.0

465
425
370
330
214
75

561
593
555
514
529
521
434

0
0
0.4
0.5
0.9
4.1
6.0
>18

17.7
17.6
25.5
26.3
30.4
29.6
31.1

0
0.01
0.01
0.03
0.05
0.18
0.18

lips

Volume fraction Vf , yield stress r0:2% determined using 0.002 strain oset method, fracture stress rf , plastic strain eplastic , impact toughness and
area fraction of shear lips Af are listed. Compressive test runs on pure La sample was stopped without reaching fracture.

600

600

500

500

400

400

300

300

200

200

100

100
0

0
0.0

0.2

0.4

0.6

0.8

critical volume fraction in the vicinity of 40% dendrites,


beyond which the compressive and tensile ductility of
the composites escalates rapidly. The plastic strain is
signicantly increased to about 56% for composites
with 50% of uniformly distributed dendrites in the
amorphous matrix; this compares quite favorably with
the 9% tensile ductility of pure La, which can likely be
considered as an upper bound on the potential ductility
in the BMG composites. As we will see subsequently, we
believe that this sudden increase in ductility near
Vf  0:40 is associated with a topological (percolation)
transition in the composites.
The Charpy impact tests showed similar general
trends to those described above for compression and
tensile tests, as shown in Fig. 6. The BMG composite
samples yielded an average impact toughness of up to
60% greater than that of monolithic BMG specimen. The
central region of each Charpy specimen was relatively

1.0

Volume Fraction of dendrites

600

500

500

400

400

300

300

200

200

100

100

0
0.0

0.2

0.4

0.6

0.8

Plastic Strain (%)

600

Compressive FractureStress (MPa)

Compressive Yield Stress (MPa)

(a)

(b)

Tensile Fracture Stress (MPa)

TensileYield Stress (MPa)

The relation between the volume fraction of dendrites


and tensile/compression plastic strain before failure is
shown in Fig. 5. These data suggest that there is a

5
4
3
2
1
0

1.0

Volume Fraction of dendrites

0.0

0.1

0.2

0.3

0.4

0.5

0.6

Volume Fraction of Dendrites


Fig. 4. (a) Tensile and (b) compression (j) yield stress, (N) fracture
stress of dendrite reinforced metallic glass matrix composites can be
reasonably described by the rule-of-mixtures. Lines are drawn to guide
the eyes.

Fig. 5. Relation of tensile (j) and compressive (N) plastic strain area
with volume fraction of dendrites in the matrix. The lines are added to
aid the eyes.

4126

M.L. Lee et al. / Acta Materialia 52 (2004) 41214131

4.2. Yield asymmetry

0.15
28
26

0.10

24
22

0.05

20
18

0.00

16
0.0

0.1

0.2

0.3

0.4

0.5

0.6

Volume Fraction of Dendrites


Fig. 6. Relation of Charpy impact toughness (j) and area fraction of
shear lip (N) with volume fraction of dendrites in the matrix. The
dashed lines are added to aid the eyes.

at corresponding to plane-strain, mode I fracture, and


was anked by clear shear lips at the edges. The higher
energy absorption during fracture was correlated with
the increasing dominance of these shear ow regions. As
shown in Fig. 6, the projected area fraction of the shear
lips increased with Vf , following much the same trend
as the tensile/compressive ductility and the impact
toughness (Fig. 5).
The above results show that although in situ reinforcements can oer considerable gains in the ductility
and toughness of BMG materials, these gains are quite
marginal below a critical value of Vf . In the present Labased system, a critical volume fraction near 40% of
dendrites is needed in the amorphous matrix to improve
the ductility of the composite signicantly. It is not clear
why the onset of the ductility enhancement should occur
so suddenly at a single value of Vf , but we believe that it
is associated with percolation of shear localization networks through the material; for high enough volume
fractions, the formation of a fully contiguous surface of
shear localization through the glass matrix could be
suppressed as the characteristic spacings of shear bands
and dendrites become similar. Since the dendrite phase
may itself be expected to form an interconnected structure at volume fractions in the range of 3040%, the
observation that macroscopic failure is suppressed in the
same range is suggestive. However, because reinforcement particles may act both as nucleation sites as well as
obstacles to shear band propagation, this is a complex
percolation problem, and the critical value of Vf will
certainly be dependent upon details of the glass composition and structure, as well as, e.g., the shape and
average length scale of the reinforcement phases. Further discussion of the fracture mechanisms of these
materials and their relationship to microstructural topology will be presented in a later section.

Recently, several research groups have been directly


examining the role of pressure- or normal stress on
plastic yield of metallic glasses [1319], and have independently concluded that yield is not solely controlled
by the maximum shear stress. Early works by Donovan
[13] and Ert
urk and Argon [20] noted some asymmetry
between the values of the yield stress for states of net
tensile and compression in Pd- and Co-based metallic
glasses, respectively. More recently, Lowhaphandu,
Lewandowski and co-workers [14,17,18] have used
combined uniaxial-plus-pressure loading to explore the
yield and failure of Zr-based metallic glasses. These
authors showed that a small pressure-dependence of the
yield stress is manifested as higher strength in compressive loading states as compared with states of net
tension. Lund and Schuh [21,22] performed molecular
simulations of multiaxial deformation in a model metallic glass to identify the form of the yield criterion and
provide an explanation for this eect. They found a
pronounced asymmetry between the magnitudes of the
yield stresses in tension and compression, with the uniaxial compression strength approximately 24% higher,
and proposed that the onset of yield may be better described by the MohrCoulomb criterion than more
typical criteria such as that of von Mises.
As shown in Table 2, we have found that there is a
measurable asymmetry between the magnitudes of the
yield stresses in tension and compression for the present
BMG composite samples. The compressive yield stresses
are generally higher than those measured in tension by
about 631%. Another important observation from the
data in Table 2, which is plotted in Fig. 7, is that with
increasing reinforcement in the amorphous matrix, the
asymmetry between the tensile and compressive yield

Asymmetry between yield stresses


in tension and compression (%)

30

Area Fraction of Shear Lip

Charpy Impact Toughness (KJ/m2)

0.20
32

35
30
25
20
15
10
5
0.35

0.40

0.45

0.50

Volume Fraction of Dendrites


Fig. 7. Amount of asymmetry between tensile and compressive yield
stresses with volume fraction of dendrites.

M.L. Lee et al. / Acta Materialia 52 (2004) 41214131

stress for specimens of the same composition seems to


decrease. Unfortunately, a 0.2% oset stress could not
be obtained for the specimens of lower volume fractions,
but in the range from Vf 0:37 to 0.53, the loss of
plastic asymmetry occurs rather suddenly. Given that
this is the same range of volume fractions over which
ductility quickly increases in these materials (Figs. 5 and
6), we propose that there is a fundamental transition
from glass-phase controlled plasticity to crystallinephase dominated yield in this range. For most crystalline
metals, the yield criterion is dominated by the maximum
shear stress, giving yield stresses of equal magnitude in
either tension or compression [2326]. The trend observed in Fig. 7 may therefore be indicative of a transition from glass-controlled yielding following an
asymmetric yield criterion (like that of Mohr and Cou-

4127

lomb) to plasticity dominated by ow of the crystalline


reinforcements with a symmetric yield criterion.
4.3. Fracture morphology
Very few data are available on tensile behavior of
BMGs [18,2731] and only Johnson et al. [6,10] have
reported tensile results for dendritic-reinforced composites. However, those authors did not provide tensile
stressstrain curves or discuss the fracture behavior of
their tensile specimens in detail. Recently, Zhang et al.
[32] proposed that the failure of a BMG material is a
competitive process depending on the loading mode and
the inhomogeneity of the microstructure, and reported
that tensile rupture, with a fracture surface inclined
hT 90 to the loading axis, occurred in an annealed

Fig. 8. Typical tensile fracture surface of (a) monolithic BMG specimen, composites with (b) Vf 0:07, (c) Vf 0:37, and (d) Vf 0:5.

4128

M.L. Lee et al. / Acta Materialia 52 (2004) 41214131

BMG with a high volume fracture of nanocrystals. The


two BMG compositions (Zr52:5 Ni14:6 Al10 Cu17:9 Ti5 and
Zr59 Cu20 Al10 Ni8 Ti3 ) with hT 90 reported in their
work were annealed for partial (Vf > 50%) or full crystallization. They explained that the tensile fracture mode
occurs in crystallized metallic glasses because of embrittlement by the intermetallic nanocrystals. In contrast, in the present BMG systems with ductile
reinforcement, we see a fracture angle near 90 for ten-

sile tests on all of the specimens. As shown in Fig. 8, not


only does the monolithic La-based BMG fracture at an
angle of 90 under tensile loading, the dendrite-reinforced composites also showed hT of 90, although with
some additional evidence of shear deformation around
the circumference of the fracture surfaces (Fig. 8).
Fractographic observation of the tensile specimens in
Fig. 8 revealed signicant dierences in fracture morphology for the composites as compared to the mono-

Fig. 9. Typical compression fracture surface of (a) monolithic BMG specimen and composites with (b) Vf 0:07, (c) Vf 0:37 and (d) Vf 0:5.

M.L. Lee et al. / Acta Materialia 52 (2004) 41214131

lithic BMG; with increasing volume fraction of dendrites in the matrix, the fracture surface transitions to a
more obvious dimpled surface characteristic of extensive
plastic ow accompanied by microvoid formation and
coalescence. What are presumed to be dendrite branches
were also observed in the dimples of the composite
fracture surfaces, particularly those with a large fraction
of reinforcement.
The compressive fracture surface of the monolithic
BMG exhibits the typical vein-like structure, as shown
in Fig. 9(a), although the failure clearly occurs in a shear
mode under compressive loads. The vein structure has
been widely attributed to melting or viscous ow of the
metallic glass [36]. In the case of the composite specimens, the compressive fracture surfaces are more complex, although failure occurs by a shear-o mechanism
in each case. Vein patterns are still observed, but have
shorter characteristic length scales as the dendrite
loading fraction increases; the fracture surface becomes
rougher with higher Vf .
For metallic glass materials obeying a MohrCoulomb yield criterion, shear planes are expected to occur
at angles dierent from 45 with respect to the stress
axis. In our compression specimens we nd that the
samples rarely exhibit a planar fracture surface as shown
in Fig. 9, so only very approximate measurements of the
compression fracture angle could be made. We nd that
the compressive fracture angle hc of monolithic Labased BMG compression sample is between about 40
and 45, which is consistent with most prior reports
[2,6,10,3335]. Similar angles were observed for all of
the composite specimens as well, but the measurements
were not accurate enough to discern an obvious trend
with Vf .
4.4. Characteristic length scales
As shown in Figs. 57, the BMG composites can be
tailored to exhibit excellent combinations of strength
and ductility. Presumably the suppression of quasibrittle behavior in the composites is related to the arresting of strain localization by the reinforcing phase. As
shown in Fig. 10, for the monolithic amorphous sample
loaded in compression, fracture occurs catastrophically
along one or a few dominant shear bands. When dendrites are incorporated, they serve as a network in the
amorphous matrix which restricts the shear banding to
comparatively isolated regions and delays localized
shear-o through the entire sample. In order for a single
surface of plastic shear ow to develop, the shear bands
have to either (i) force a shear distortion through the
ductile dendrites or (ii) serrate to bypass the dendrites.
In both ways, more energy is required to force global
plastic failure on a single shear surface, and this leads to
improved compression ductility of the composites. At
the same time, plastic ow and work hardening of the

4129

Fig. 10. SEM images showing shear bands on the side of compression
monolithic BMG specimens. Arrows show the direction stress applied
during compression.

dendrites through dislocation motion leads to global


work hardening behavior in the composites.
As described above, we believe that the very rapid
increase in ductility near Vf  0:4 is associated with the
topological suppression of shear surfaces that span the
specimen, i.e., a percolation transition. Our experimental observations also support this hypothesis in that
there is a cross-over of length scales near the critical
volume fraction. Specically, we note that the shear
bands on the side of the monolithic BMG samples (see
Figs. 10 and 11(a)) have a characteristic average spacing
in the range of 520 lm. In the most dilute composites
with Vf 0:07, the inter-dendritic spacing is on the order of 1040 lm. It seems reasonable that with an interdendritic spacing generally larger than the natural
shear band spacing, these dendrites need not substantially aect the path of the propagating shear bands.
This is indeed the case, as these composites failed catastrophically without obvious ductility (Figs. 5 and 6). In
contrast, in the composites with Vf  0:4, the typical
inter-dendritic spacing in a 2-D section is in the range of
28 lm, generally smaller than the natural shear band
spacing in the monolithic BMG. In this case there is a
substantially stronger interaction between shear bands
and dendrites, leading to strain hardening and limited
ductility. Beyond the critical volume fraction, the ductility of the composites is signicantly improved; as the
interdendritic spacing approaches zero above Vf  0:4, a
continuous network of dendrites begins to appear, and
shear bands must be accommodated by plastic ow of
the crystalline phase.
A similar observation can also be made on the side of
the fractured tensile specimens, as shown in Fig. 11.
However, due to the dierent mode of fracture of these
materials under tensile loading, microcracks or voids
can be seen on the side of the fractured samples instead

4130

M.L. Lee et al. / Acta Materialia 52 (2004) 41214131

Fig. 11. SEM images showing interaction of microcracks with dendrites on the side of tensile specimens. (a) Monolithic BMG, (b) composite with
Vf 0:37, (c) composite with Vf 0:5 and (d) higher magnication of (c). Arrows show the direction of stress under tensile testing.

of shear bands. When the dendrites are more closely


spaced there is a ner dispersion of microcracks
(Fig. 11(d)). In fact, we occasionally observe that the
microcracks traverse dendrite particles. This result again
emphasizes the importance of microstructural topology
to the deformation and ductility of in situ BMG composites. It seems quite fortunate that dendritic phases
can help distribute both shear bands and tensile microcracks; judging from the ductility gains seen in Fig. 5,
the in situ reinforcements appear equally potent against
either mode of strain localization.

5. Conclusions
A systematic study has been carried out to identify
the eect of second phase reinforcements on the mechanical properties of amorphous alloys based on the
composition La86y Al14 (Cu, Ni)y (y 124). The most
salient results of this work are as follows:
First, we observe that like most composites, these
BMG-matrix composites exhibit tensile and compressive yield strengths that generally obey a rule-of-mixtures relationship. Their ductility and impact
toughness however exhibit distinctly non-linear behavior, with considerably more plastic ow occurring
for reinforcement volume fractions exceeding a critical value near 40%. We believe that this non-linearity
is associated with a topological change related to the
disruption of shear surfaces that cross the specimen.
This percolation-based explanation is also consistent

with an observed cross-over of the shear band and inter-dendritic spacings as the volume fraction of reinforcement exceeds about 40%.
Second, where prior studies have suggested that dendrites help distribute shear bands through a metallic
glass, we also nd that tensile microcracks are dispersed in a similar fashion by in situ reinforcements.
As a consequence, both tensile and compressive ductility can be substantially enhanced by forming composites, though the fracture mode is dierent in the
two cases.
Finally, we have also observed an asymmetry between the magnitudes of the yield stresses in tension
and compression in these BMG-based materials, in
line with prior work on amorphous metals. However,
we also observe for the rst time that with increasing
volume fraction of dendrites in the matrix, this asymmetry becomes smaller. This result is consistent with
a transition from MohrCoulomb-type yielding characteristic of metallic glasses, to symmetric yielding
commonly observed in metal crystals.

Acknowledgements
This work was supported primarily by the SingaporeMIT Alliance. C.A.S. acknowledges the support of the
US Army research oce, under contract DAAD19-031-0235, although the views expressed in this work are
not endorsed by the sponsor. Collaboration with Prof.
W. C. Carter of MIT is gratefully acknowledged.

M.L. Lee et al. / Acta Materialia 52 (2004) 41214131

References
[1] Choi-Yim H, Johnson WL. Appl Phys Lett 1997;71:3808.
[2] Choi-Yim H, Busch R, K
oster U, Johnson WL. Acta Mater
1999;47(8):2455.
[3] Conner RD, Dandliker RB, Johnson WL. Acta Mater
1998;46:6089.
[4] Das J, Loser W, Kuhn U, Eckert J, Roy SK, Schultz L. Appl Phys
Lett 2003;82:4690.
[5] Fan C, Ott RT, Hufnagel TC. Appl Phys Lett 2002;81:1020.
[6] Hays CC, Kim CP, Johnson WL. Phys Rev Lett 2000;84:2901.
[7] Hu X, Ng SC, Feng YP, Li Y. Acta Mater 2003;51:561.
[8] Kuhn U, Eckert J, Mattern N, Schultz L. Appl Phys Lett
2002;80:2478.
[9] Pekarskaya E, Kim CP, Johnson WL. J Mater Res 2001;16:2513.
[10] Szuecs F, Kim CP, Johnson WL. Acta Mater 2001;49:1507.
[11] Tan H, Zhang Y, Li Y. Intermetallics 2002;10:1203.
[12] Tan H, Zhang Y, Feng YP, Li Y. Philos Mag 2004;84:53.
[13] Donovan PE. Acta Metall 1989;37:445.
[14] Lewandowski JJ, Lowhaphandu P. Philos Mag A 2002;82:3427.
[15] Davis LA, Kavesh SJ. Mater Sci 1975;10:453.
[16] Flores KM, Dauskardt RH. Acta Mater 2001;49:2527.
[17] Lowhaphandu P, Montgomery SL, Lewandowski JJ. Scripta
Mater 1999;41:19.
[18] Lowhaphandu P, Ludrosky LA, Montgomery SL, Lewandowski
JJ. Intermetallics 2000;8:487.
[19] Vaidyanathan R, Dao M, Ravishandran G, Suresh S. Acta Mater
2001;49:3781.

[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]

4131

Erturk T, Argon AS. J Mater Sci 1987;22:1365.


Schuh CA, Lund AC. Nature Mater 2003;2:449.
Lund AC, Schuh CA. Acta Mater 2003;51:5399.
Kimura H, Masumoto T. In: Luborsky FE, editor. Amorphous
metallic alloys. London: Butterworth; 1983. p. 187.
Bulatov VV, Richmond O, Glazov MV. Acta Mater
1999;47:3507.
Courtney TH. Mechanical behavior of materials. New York:
McGraw-Hill; 1990.
Bruck HA, Christman T, Rosakis AJ, Johnson WL. Scripta
Metall 1994;30:429.
Alpas AT, Edwards L, Reid CN. Metall Trans A 1989;20:1395.
He G, Lu J, Bian Z, Chen D, Chen G, Tu G, Chen G. Mater
Trans 2001;42:356.
Inoue A, Kimura HM, Zhang T. Mater Sci Eng A 2000;294
296:727.
Inoue A, Zhang W, Zhang T, Kurosaka K. Acta Mater
2001;49:2645.
Liu CT, Heatherly L, Easton DS, Carmichael CA, Schneibel JH,
Chen CH, et al. Metall Mater Trans A 1998;29:1811.
Zhang ZF, He G, Eckert J, Schultz L. Phys Rev Lett
2003;91(4):045505.
Heilmaier MJ. Mater Proc Tech 2001;117:374.
Xing LQ, Herlach DM, Cornet M, Bertrand C, Dallas J-P, Trichet
M-F, et al. Mater Sci Eng A 1997;226228:874.
Conner RD, Choi-Yim H, Johnson WL. J Mater Res
1999;14(8):3292.
Kawamura Y, Mano H, Inoue A. Scripta Mater 2000;43:1119.

Potrebbero piacerti anche