Você está na página 1de 12

Int. J. Nanoparticles, Vol. X, No.

Y, xxxx 1

Mechanically induced cyclic glassy-nanocrystalline


phase transformations of tetragonal Zr2Ni alloys

M. Sherif El-Eskandarany
Center of Nanotechnology, Director General’s Office,
Kuwait Institute for Scientific Research,
P.O. Box 24885 Kuwait Safat 13109, State of Kuwait
Fax: +(965) 24836608
E-mail: msherif@kisr.edu.kw

Abstract: Cyclic crystalline-glassy-crystalline phase transformations have been


investigated during high-energy ball milling of tetragonal-Zr2Ni alloy powders,
using mechanical disordering (MD) method. The results show that the
metastable fcc-Zr2Ni alloy powders, which is obtained after 173 ks of MD time,
transforms into a new metastable phase of nanocrystalline big cube-Zr2Ni upon
milling for 360 ks. The lattice parameters, a0 for the metastable-fcc and big
cube phases were calculated and found to be 0.451 nm and 1.226 nm,
respectively. The obtained big-cube phase subsequently transforms to a single
glassy phase after 540 ks of MD time. This obtained glassy phase, however,
devitrificated into the same metastable big-cube phase upon increasing the MD
time to 720 ks. As the MD time increases (806 ks to 900 ks), the obtained
nanocrystalline big-cube phase transforms into the pervious metastable
fcc-Zr2Ni phase, indicating a cyclic-phase transformations between the three
obtained metastable phases. Such cyclic phase transformations are observed
several times during the milling procedure. On the basis of our results, the
destabilising effect of the defects created by the milling media (balls) is
responsible for such mechanically induced solid-state devitrification (MISSD)
that leads to the cyclic transformations.

Keywords: mechanical alloying; MA; high-energy ball milling; lattice defects;


microstructure; crystal structure; phase transformations; thermal stability;
metallic glasses; big-cube phase; metastable phase.

Reference to this paper should be made as follows: El-Eskandarany, M.S.


(xxxx) ‘Mechanically induced cyclic glassy-nanocrystalline phase
transformations of tetragonal Zr2Ni alloys’, Int. J. Nanoparticles, Vol. X,
No. Y, pp.000–000.

Biographical notes: M. Sherif El-Eskandarany is a Visiting Professor of


Advanced Materials and Nanotechnology at IMR, Tohoku University, Japan
and is working now as a Senior Research Scientist at Kuwait Institute for
Scientific Research. He obtained his Master and Doctorate degrees from
Tohoku University, Japan. He was a former Vice-President of the Academy of
Scientific Research and Technology in Egypt and one of the nanotechnology
founders in Egypt. He has published 120 peer-reviewed international journals,
130 international conference articles and three international books. He received
the Golden Medal of Excellence in Scientific Research from H.E. the President
of Egypt and the Golden Medal of Technology from JSPM, Japan.

Copyright © 200x Inderscience Enterprises Ltd.


2 M.S. El-Eskandarany

1 Introduction

Metallic glasses (Duwez, 1960) have attracted many materials scientists from different
international schools (Greer and Ma, 2007; Kato et al., 2008; El-Eskandarany and Inoue,
2006a) due to their unique properties (Yavari et al., 2007) that are not found when the
same alloys are in their stable crystalline states. Hence, metallic glasses have been
considered as promising candidates for several technical applications (El-Eskandarany,
2001; Inoue and Nishiyama, 2007). However, glassy alloys can be obtained by several
ways of fabrications (Inoue, 1999), mechanical alloying (MA) method, using ball-milling
technique has been considered as a powerful tool (El-Eskandarany and Inoue, 2006b) for
the fabrication of wide varieties of glassy powders (e.g., Bai et al., 2001; El-Eskandarany
et al., 2002; Huang et al. 2008), including those systems that cannot be synthesised by the
conventional melting and casting techniques (El-Eskandarany et al., 2003). Consolidation
of these powders within their supercooled liquid regions leads to the formation of fully
dense objects (El-Eskandarany and Inoue, 2006a, 2006b).
Zr-Ni binary system is the best known metal-metal glassy former, which can be made
amorphous by several techniques over a wide range of composition; extensive studies
have been made and reported on the crystallisation and kinetics of this system. The
present study has been undertaken to investigate the structural changes that take place
upon room temperature mechanical disordering (MD) (El-Eskandarany et al., 1991)
crystalline powders of tetragonal-Zr2Ni, using high energy ball mill. The stability of the
obtained nano-metastable powders against the shear and impact forces that are generated
by the milling media and the subsequent devitrifications was demonstrated by subjecting
the powders to longer MD times. Thus, the possibility of mechanically-induced
solid-state devitrifications (MISSD) could be investigated.

2 Experimental procedure

An ingot of Zr67Ni33, obtained by arc-melting a mixture of Zr (99.9 at. %) and Ni


(99.9 at. %) metals, was crushed down into particles, using a tempered chrome steel
mortar and pestle. The ground particles were then screened to obtain finer powders
(50 μm to 60 μm in diameter) and used as the starting materials for the MD process. The
starting powders were charged into tempered chrome steel vials (250 ml in volume) and
sealed together with 50 tempered chrome steel balls (10 mm in diameter) in an argon
atmosphere glove box. The ball-to-powder weight ratio was maintained as 14:1. The MD
process was performed in a planetary ball mill (Fritsch P5) at a rotation speed of 2.1 s–1.
In order to avoid an increase in the vial temperature, the milling procedure was
periodically interrupted every 1.8 ks and then halted for 1.8 ks under a continuous flow of
air. The MD experiments were performed two or three times under the same experimental
conditions to confirm the reproducibility of the results. The ball milling was stopped after
selected milling times and a small amount (about 500 mg) of powders were taken from
the vials in the glove box. The as-milled samples were characterised by means of X-ray
diffraction (XRD) employing CuKα radiation, transmission electron microscopy (TEM),
using a 300 kV field emission microscope, scanning electron microscopy (SEM) using
15 kV field emission electron microscope, differential scanning calorimetry (DSC) at a
constant heating rate of 0.67 K/s and under flow of a purified argon gas. High resolution
transmission electron microscope/energy dispersive spectroscopy (HRTEM/EDS), using
Transformations of tetragonal Zr2Ni alloys 3

an electron beam of 5 nm has been employed for analysing the concentration of Zr and
Ni, and the degree of iron contamination in the milled powders. In addition, the oxygen
content was determined by the helium carrier fusion-thermal conductivity method.
Annealing the powder samples was performed at the desired temperatures in sealed
copper DSC cells under continuous flow (2.5 ml.s–1) of purified argon (99.999 wt. %)
flow. As soon as the annealing procedure is achieved, the DSC was rapidly cooled and
the samples were subsequently removed from the DSC heating chamber. The as-annealed
samples were structurally analysed by means of XRD and TEM/EDS techniques.

3 Results and discussions

The XRD patterns of the ball-milled powders after selected MD times are displayed
together in Figure 1.

Figure 1 XRD patterns of ball-milled Zr2Ni powders after, (a) 0 ks (b) 173 ks (c) 360 ks
(d) 540 ks of MD (e) 720 ks (f) 806 ks (g) 900 ks of MD time

The sample at the starting stage of MD (0 ks) is a single phase of polycrystalline


tetragonal-Zr2Ni and does not show the existence of any other phases, indicated by the
presence of sharp indexed Bragg peaks of the starting ordered phase [Figure 1(a)]. These
Bragg peaks are no longer appeared in the sample that milled for
173 ks of MD and a new set of broad Bragg-lines corresponding to fcc structure,
which come from the reflections of (111), (200), (220) and (311) planes, appeared in
Figure 1(b). The analysis of the Bragg peaks for this new obtained phase confirmed the
formation of a metastable fcc-Zr2Ni (fcc-Zr2Cu structure, Mann et al., 2002) and its
lattice parameter, a0, was calculated and found to be 0.451 nm. This obtained fcc-Zr2Ni
4 M.S. El-Eskandarany

metastable phase is not capable enough to withstand against the mechanical


imperfections that came from the milling media and surprisingly transformed into a new
crystalline phase after 360 ks of MD time [Figure 1(c)]. Careful analysis of the diffracted
peaks in Figure 1(c) suggests that the powders at this stage of MD reveal polycrystalline
structure corresponding to Zr2Ni structure (E93 structure, space group Fd3m). The lattice
constant, a0 of this crystalline phase was calculated to be 1.226 nm.
When the MD time is increased to 540 ks, all of the Bragg peaks corresponding to the
formed big-cube phase have already disappeared and clear broad diffuse haloes appear,
implying the formation of a single amorphous phase with no indication of precipitation of
any residual crystalline phases [Figure 1(d)]. The image of the HRTEM of the sample
after 540 ks of MD time is shown in Figure 2(a) together with the corresponding selected
area diffraction pattern, SADP [Figure 2(b)]. Overall, the sample, which appears
featureless and homogeneous in its internal structure, shows a maze contrast with no
indication of precipitations of any crystalline phases [Figure 2(a)], implying the
homogeneity in structure within the nanoscale Moreover, the SADP displays a typical
halo-diffraction (diffuse halo) of an amorphous phase [Figure 2(b)]. The absence of any
crystalline contrast [Figure 2(a)] and the hardly vision of sharp rings or spots [Figure
2(b)] would indicate the formation of fully amorphous phase. The EDS technique was
employed to ensure the local composition of the obtained sample at this stage of MD.
These analyses (not presented here) that taken from the different regions of the milled
powder do not show any serious compositional gradient and the average composition is
being very close to the starting nominal composition of Zr67Ni33. Although the XRD
pattern of the sample that milled for 576 ks (not presented here) indicates that the
obtained powder maintains its unique random structure, as far as indicated by broad
primary and secondary diffraction maxima, the HRTEM image of this sample [Figure
2(c)] shows obvious nano fringes of local ordered regions with different contrast (see, for
example, the indicated arrows), which are embedded in the amorphous matrix.
Furthermore, the nano beam diffraction pattern (NBDP) that taken from Region IX in
Figure 2(c) shows a halo-diffraction pattern coexisted with several spots [Figure 2(d)],
implying the precipitation of a crystalline phase in the residual amorphous matrix. The
EDS analytical results (not shown here) of the Regions IX and X could not detect any
sharp compositional fluctuations between those selected neighbour regions, implying the
homogeneity of the obtained product at this stage of MD.
Transformations of tetragonal Zr2Ni alloys 5

Figure 2 HRTEM images (a, c) and the corresponding SADP (b, d) of ball-milled Zr2Ni powders
after 540 ks (a, b) and 576 ks of the MD time

Notes: The indexed circles in images (a, c) refer to the EDS examined regions. The
fringes, which are embedded in the glassy matrix in image (c) belong to the
big-cube Zr2Ni metastable phase.
The XRD pattern of the sample that was obtained after continuous ball milling for 720 ks
of MD time is displayed in Figure 1(e). Obviously, the amorphous phase is devitrificated
into a nanocrystalline phase, characterised by the disappearance of the halo-diffraction
peaks and the appearance of Bragg lines of a crystalline phase. Surprising came from
when the analysis of the diffracted lines in Figure 1(e) indicates that the yielded
crystalline phase is a big-cube Zr2Ni that has a0 similar to that one, which was previously
formed after 360 ks of MD time [Figure 1(b)]. The formed metastable big cube-phase of
this stage of milling consists of spherical nanocrystalline grains ranging from 20 nm to
75 nm in diameter, as presented in the dark field image (DFI) of Figure 3(a). The
appearance of sharp-ring fashion at the corresponding SADP [Figure 3(b)] with the
absence of spot-like patterns suggests the formation of fine-grain structure and the
disappearance of the previously formed amorphous phase. It is worth noting that the local
EDS analyses that taken from several regions in Figure 3(a) does not show any surprising
compositional differences, indicating that the MISSD takes place in a typical
polymorphous reaction.
6 M.S. El-Eskandarany

Figure 3 DFI (a) and the corresponding SADP (b), and BFI (c) and its SADP of ball-milled
Zr2Ni powders after ball milling for 720 ks (a, b) and 900 ks, respectively

Note: The indexed circles in the images refer to the EDS examined regions.
The stability of the yielded big-cube phase against the mechanical deformation during
ball milling procedure has been realised so that the powders were furthered milled to
806 ks. As it is expected, this phase was transformed to a different metastable phase,
indicated by the shown Bragg peaks in Figure 1(f), which corresponding to an fcc-Zr2Ni
metastable phase. This fcc-metastable phase has almost the same a0 of that phase,
which obtained after milling for 173 ks [see Figure 1(b)]. This would suggest
cyclic-metastable-metastable-phase transformations, which was firstly demonstrated by
El-Eskandarany et al. (1997). Increasing the MD time to 900 ks does not lead to any
further phase transformations and the end-product still maintains its fcc-structure, as
displayed in the XRD pattern in Figure 1(g). The obvious broadening in the Bragg lines
of this metastable phase indicates the formation of a nanocrystalline phase. More precise
information about the local structure and composition of the powder of the end-product
(900 ks) powders could be obtained by TEM/EDS analyses. The bright field image (BFI)
of the powders at this stage of milling indicates the formation of nanomaterials that do
not show any dramatic changes in composition and have uniform nano-structure with
spherical grains of less than 10 nm in diameter [Figure 3(c)]. In addition, the indexed
NBDP of a selected region indicates the presence of a single fcc phase in the milled
powders with the absence of any other metastable phases (amorphous and/or big-cube
phases).
The thermal stabilities of the three metastable products (glassy, big-cube and
fcc-Zr2Ni alloys) that are yielded cyclically during the ball milling process were
examined by DSC. The DSC thermograms of these products are displayed in Figure 4
after selected MD times. The DSC scan of the sample that obtained after 173 ks of MD
time reveals a single exothermic peak at 941 K [Figure 4(a)]. The X-ray analysis of the
sample taken after this DSC run shows the formation of tetragonal-Zr2Ni phase. The
enthalpy change of crystallisation, ΔH xfcc→ordered is calculated from the area under the
Transformations of tetragonal Zr2Ni alloys 7

exothermic peak in Figure 4(a) and found to be –2.28 kJ/mol. ΔH xfcc→ordered here is
considered as the heat of formation of tetragonal-Zr2Ni alloy.

Figure 4 DSC thermograms of ball-milled Zr2Ni powders after, (a) 173 ks (b) 360 ks (c) 432 ks
(d) 470 ks (e) 540 ks (f) 576 ks (g) 630 ks (h) 720 ks (k) 900 ks of MD time

Note: The XRD patterns of the annealed samples at I, II, III and IV are shown in Figure 5
(a, b, c and d, respectively).
However, this exothermic reaction still appears in the DSC curve of the
sample that milled for 360 ks of MD time [Figure 4(b)], the general feature
of the thermograms is changed, characterised by the appearance of a
low-temperature broad exothermic reaction that took place at 806 K. The XRD
pattern of the sample that taken after the first exothermic reaction (850 K)
shows the formation of metastable fcc-Zr2Ni phase. The enthalpy change of
crystallisation, ΔH xbig −cube→ fcc that realised due to this exothermic reaction is calculated
and found to be –0.86 kJ/mol. Here, the ΔH xbig −cube→ fcc is considered as the heat of
formation of metastable fcc-Zr2Ni alloy. In this DSC scan [Figure 4(b)] the area under the
second exothermic reaction, which is considered as ΔH xfcc→ordered , is calculated
to be –2.13 kJ/mol. Thus, the total enthalpy change of crystallisation,
ΔH xbig cube→fcc→ordered (Δ H xbig cube→fcc→ordered = Δ H xbig cube→ fcc + Δ H xfcc→ordered ) is –2.99
kJ/mol.
Both peaks tend to disappear and becoming hardly visible when the MD time
increases to 432 ks [Figure 4(c)], whereas new peaks appear at this DSC scan. However,
the first reaction, which is an endothermic type and refers to the glass transition
temperature (Tg) of the formed glassy powders can be seen hardly at 691 K, the second
peak, which is an exothermic reaction and results due to the crystallisation (Tx) of the
glassy phase is obvious and takes place at 736 K (onset temperature), as shown in Figure
4(c). These two opposite reactions became more pronounced for those samples that
obtained after 470 ks [Figure 4(d)] and 540 ks [Figure 4(e)] of MD time. This implies an
increase in the volume fraction of the formed glassy phase. The X-ray analysis of the
sample that obtained after completing this DSC run (750 K) shows the formation of
8 M.S. El-Eskandarany

tertragonal-Zr2Ni phase, suggesting that a glassy-ordered phase transformation at the


observed Tx. It is worth noting that the onset temperature of both reactions does not
change markedly, suggesting the absence of any compositional changes during the MD
procedure. The wide value of the supercooled liquid region, ΔTx (ΔTx, equal to Tx – Tg at
46 K) for a binary metallic system, suggests that the ball-milled Zr2Ni alloy has a good
glass-forming ability. Moreover, the enthalpy change of crystallisation, ΔH xglassy→ordered
of the glassy phase that is calculated from the area under the exothermic reaction in
Figure 4(e) and is found to be –9.2 kJ/mol.
The DSC scan for the sample that obtained after 576 ks of MD time [Figure 4(f)]
shows a different feature in respect to the intensity of the exothermic reaction that
becoming weak. In addition, the crystallisation process takes place throughout several
stages, indicated by a series of exothermic reactions, which are grown and becoming
more visible for the sample that milled for 630 ks of MD time [Figure 4(g)]. All of these
events suggest a drastic decreasing of the formed glassy phase and a grown of new
phase(s). It is worth mentioned that the thermal characterisation of the residual glassy
phase, characterised by Tg, Tx and ΔTx does not change, implying the absence of any
compositional changes. Contrary to the first exothermic reaction that appeared in the scan
of Figure 4(g), the second exothermic peak is taking place widely at 756 K. The last
reaction of this DSC scan, however, takes place at higher temperature, as high as 850 K.
In order to explore the origin of the above mentioned reactions, four individual
heating runs were achieved for the sample, which was milled for 540 ks of MD time and
each sample was checked carefully by X-ray analysis. These samples [#I, #II, #III and
#IV, Figure 4(g)] were individually post-annealed at temperatures 662 K, 737 K, 790 K
and 987 K, respectively for 300 s in the DSC under flow of pure argon.
The sample (#I), which was annealed at 662 K maintains its original characterisations
of having a random structure without any devitrifications to any metastable phase(s),
suggested by the clear broad diffused halos that are visible in Figure 5(a). Annealing of
this glassy phase at 737 K [#II in Figure 4(g)] (after its crystallisation peak) leads to a
devitrification into a metastable big-cube phase [Figure 5(b)], similar to that one obtained
after 360 ks and 720 ks of MD times [see Figures 1(c) and 1(e), respectively].
This formed metastable big-cube Zr2Ni phase does not maintain its unique crystalline
structure and transformed into another metastable phase of fcc-Zr2Ni upon post-annealing
the powders at 790 K [#III in Figure 4(g)], as was confirmed by the XRD pattern in
Figure 5(c). It is worth noting that this formed metastable fcc-phase has almost the same
a0 of those products, which yielded after 173 ks and 806 ks of MD times [Figures 1(b)
and 1(f), respectively]. The XRD pattern of the sample powder [#IV in Figure 4(g)],
which was annealed after achieving the last exothermic peak (987 K), indicates that this
exothermic reaction takes place due to transformation of the metastable fcc-Zr2Ni into a
the most-stable phase of tetragonal-Zr2Ni [Figure 5(d)], in good agreement with that
XRD pattern of the starting (before milling) powders of Zr2Ni [Figure 1(a)].
Transformations of tetragonal Zr2Ni alloys 9

Figure 5 The XRD patterns of ball-milled Zr2Ni powders after ball milling for 630 ks of MD
time plus annealing at, (a) 662 K (b) 737 K (c) 790 K (d) 987 K

Note: The annealing procedure was performed in a DSC under purified (99.9995 wt.%)
Ar gas flow for 180s.
The area under the first exothermic reaction in Figure 4(g) presents the heat formation of
big-cube Zr2Ni phase upon annealing the Zr2Ni glassy phase that is obtained after 630 ks
of MD time (Δ H xglassy →big cube or Δ H big
for
cube ), and found to be –5.89 kJ/mol. The

ΔH xbig cube→ fcc or the heat formation of metastable fcc-Zr2Ni alloy [ ΔH fcc
for
, the
calculated area under the second exothermic peak in Figure 4(g)] is found to be
–0.93 kJ/mol. The ΔH xfcc→ordered [the area under that last exothermic peak in Figure
4(h)] or so called the heat formation of ordered-Zr2Ni alloy ΔH ordered
for
is calculated and
found to be –1.74 kJ/mol.
Thus, the total heat formation of tetragonal Zr2Ni alloy can be calculated as;
ΔH ordered
for
(ΔH xglassy→big cube→fcc→ordered ) = ΔH xglassy→big cube + ΔH xbig cube→ fcc + ΔH xfcc→ordered .
Here, ΔH ordered
for
, which is calculated to be –8.65 kJ/mol has a very close value to the
ΔH ordered
for
(–9.2 kJ/mol), which obtained upon single phase transformation of
glassy-ordered phase transformation for the sample that milled for 540 ks of MD time.
10 M.S. El-Eskandarany

The DSC scan of the powders, which were milled for 720 ks (originally metastable
big-cube phase) is displayed in Figure 4(h). One can notify the disappearance of the
endothermic reaction, which indicates the absence of the glassy phase [Figure 4(h)]. The
disappearance of the glassy phase is also indicated by the absence of the sharp
exothermic reaction, which was replaced by a broad exothermic peak that took place at
731 K [Figure 4(h)]. At a temperature well above this broad exothermic peak (832 K),
this big-cube Zr2Ni phase transformed into a more stable phase of fcc-Zr2Ni, as examined
by the X-ray analysis.
Finally, we have examined the thermal stability of the metastable fcc-Zr2Ni powders
that were obtained after 900 ks of MD time. The DSC curve of this alloy is shown in
Figure 4(k). Whereas the second exothermic peak in Figure 4(g) was vanished, the
high-temperature exothermic reaction still appears at a peak temperature of 900 K. The
XRD pattern of the sample, which was taken after annealing the powders at 985 K
for 300 s (not shown here), implies that this exothermic reaction takes place by
the transformation of the metastable fcc-Zr2Ni into the most-stable phases of
tetragonal-Zr2Ni.
In order to explore the influence of contaminations on the phase transformations of
Zr2Ni powders during the different stages of MD, the iron and oxygen contents in several
milled samples were chemically analysed and plotted in Figure 6 as a function of MD
times. The iron, which comes from the stainless steel milling tools, monotonically
increased with increasing MD time and saturated at a level of 0.57 at.% after 540 ks of
MD time. Likewise, the oxygen contamination increases significantly during the first
kiloseconds of milling to a value of 0.19 at.% after 86 ks of the MA time. At the final
stage of milling (173 ks to 900 ks), the oxygen content saturates at a value of 0.24 at.%.

Figure 6 Effect of MD time, MD time on iron and oxygen contamination contents in the
ball-milled Zr2Ni powders
Transformations of tetragonal Zr2Ni alloys 11

4 Conclusions

Our results demonstrated cyclic metastable phase transformations during MD of


tetragonal-Zr2Ni intermetallic powders, using high-energy ball mill. After 173 ks of
milling time, an fcc-Zr2Ni metastable phase is obtained and transformed into
ordered-Zr2Ni upon heating to 941 K. The enthalpy change of transformation,
ΔH fcc→ordered was found to be –2.28 kJ/mol. This fcc phase transforms into another
metastable phase (big-cube Zr2Ni) upon milling to 360 ks, using the same milling
conditions. This big-cube Zr2Ni transforms to fcc-Zr2Ni at 850 K with ΔH big −cube→ fcc of
–0.86 kJ/mol. A single glassy phase of Zr2Ni is obtained as a result of further milling for
540 ks. This glassy phase that has Tg, Tx and ∆Tx of 691 K, 736 K and 41 K, respectively,
transforms into ordered-Zr2Ni with ΔH xglassy→ordered of –9.2 kJ/mol. Increasing the ball
milling time (720 ks) leads to the devitrification of the obtained glassy phase and the
formation of nanocrystalline big-cube Zr2Ni powders. This big-cube phase is no longer in
the milled powders, when the milling time increased to 806 ks, and transformed into the
same metastable phase of fcc-Zr2Ni that was early obtained after 173 ks of the time. We
have attributed these transformations, which took place in the ball mill to the disability of
the formed metastable phases to withstand against the impact and shear forces that are
generated by the milling media (balls). When the milling time is increased, the formed
metastable phase is subjected to crystalline defects (points and lattice defects) that rise or
lowers the free energy of the formed phases, being dependant on the milling time. Our
results cannot be explained by the concept of the local temperature rise during milling
because the obtained metastable phases are periodically transformed to each other upon
further milling time and under the same milling conditions.

References
Bai, H.Y., Michaelsen, C., Gente, C. and Bormann, R. (2001) ‘Amorphization by mechanical
alloying in metallic systems with positive Gibbs energy of formation’, Phys. Rev. B, Vol. 63,
pp.064202–064211.
Duwez, P. (1960) ‘Non-crystalline structure in solidified gold–silicon alloys’, Nature, Vol. 187,
pp.869–870.
El-Eskandarany, M.S. (2001) Mechanical Alloying for Fabrication of Advanced Engineering
Materials, 1st ed., William Andrew Publishing, New York, USA.
El-Eskandarany, M.S. and Inoue, A. (2006a) ‘Synthesis of new bulk metallic glassy
Ti60Al15Cu10W10Ni5 alloy by hot-pressing the mechanically alloyed powders at the
supercooled liquid region’, Met. Trans. A, Vol. 37, pp.2231–2238.
El-Eskandarany, M.S. and Inoue, A. (2006b) ‘Hot pressing and characterizations of mechanically
alloyed Zr52Al6Ni8Cu14W20 glassy powders’, J. Mater. Res., Vol. 21, pp.976–987.
El-Eskandarany, M.S., Aoki, K. and Suzuki, K. (1991) ‘Difference between mechanical alloying
and mechanical disordering in the amorphization reaction of Al50Ta50 in a rod mill’, J. Alloys
Compounds, Vol. 177, pp.229–244.
El-Eskandarany, M.S., Aoki, K., Sumiyama, K. and Suzuki, K. (1997) ‘Cyclic
crystalline-amorphous transformations of mechanically alloyed Co75Ti25’, Appl. Phys. Let.,
Vol. 70, pp.1679–1681.
El-Eskandarany, M.S., Saida, J. and Inoue, A. (2002) ‘Amorphization and crystallization behaviors
of glassy Zr70Pd30 alloys prepared by different techniques’, Acta Mater., Vol. 50, pp.2725–
2736.
12 M.S. El-Eskandarany

El-Eskandarany, M.S., Zhang, W. and Inoue, A. (2003) ‘Mechanically induced solid-state reaction
for synthesizing new multicomponent Ta55Zr10Ni10Al10Cu15 glassy alloy powders with
extremely wide supercooled liquid region’, J. Alloys Comp., Vol. 350, pp.222–232.
Greer, A.L. and Ma, E. (2007) ‘Bulk metallic glasses: at the cutting edge of metals research’, MRS
Bulletin, Vol. 32, pp.611–615.
Huang, L.J. et al. (2008) ‘Crystallization kinetics of Mg65Cu25Nd10 amorphous alloy’,
J. of Non-Crystalline Solids, Vol. 354, pp.1048–1053.
Inoue, A. (1999) Bulk Amorphous Alloys: Practical Characteristics and Applications, in Magini,
M. and Wöhlbier, F.H. (Eds.), 1st ed., Trans Tech Publications, Switzerland.
Inoue, A. and Nishiyama, N. (2007) ‘New bulk metallic glasses for applications as
magnetic-sensing, chemical, and structural materials’, MRS Bulletin, Vol. 32, pp.651–658.
Kato, H., Chen, H.S. and Inoue, A. (2008) ‘Relationship between thermal expansion coefficient and
glass transition temperature in metallic glasses’, Scripta Materialia, Vol. 58, pp.1106–1109.
Mann, I.P., Chattopadhyay, P., Banhart, F. and Fecht, H.J. (2002) ‘Formation of
face-centered-cubic zirconium by mechanical attrition’, Appl. Phys. Let., Vol. 81,
pp.4136–4139.
Yavari, A.R., Lewandowski, J.J. and Eckert, J. (2007) ‘Mechanical properties of bulk metallic
glasses’, MRS Bulletin, Vol. 32, pp.635–638.

Você também pode gostar