Anda di halaman 1dari 14

Journal of Alloys and Compounds 350 (2003) 232–245 L

www.elsevier.com / locate / jallcom

Glass-forming ability and magnetic properties of mechanically solid-state


reacted Co 1002x Ti x alloy powders
a, a b
M. Sherif El-Eskandarany *, Wei Zhang , A. Inoue
a
Inoue Superliquid Glass Project, Exploratory Research for Advanced Technology, Japanese Science and Technology Corporation, Yagiyama-minami,
2 -1 -1, Sendai 982 -0807, Japan
b
Institute for Materials Research, Tohoku University, Katahira 2 -1 -1, Sendai 980 -8577, Japan
Received 19 June 2002; received in revised form 12 July 2002; accepted 12 July 2002

Abstract

A high-energy ball milling technique using the mechanical alloying method has been employed for fabrication of glassy Co 1002x Tix
(25#x#67) alloy powders at room temperature. The fabricated glassy alloys in the Co-rich (33$x) side exhibit good soft magnetic
properties. The binary glassy alloys for which the glass transition temperatures (T g ) have rather high temperatures (above 800 K), show
large supercooled liquid regions before crystallization (DT x larger than 50 K). The reduced glass transition temperature (ratio between T g
and liquidus temperatures, T l (T g /T l )) was found to be larger than 0.56. We have also performed post-annealing experiments on the
mechanically deformed Co / Ti multilayered composite powders. The results show that annealing of the powders at 710 K leads to the
formation of a glassy phase (thermally enhanced glass formation reaction), of which the heat of formation was measured directly. The
similarity in the crystallization and magnetization behaviors between the two classes of as-annealed and as-mechanically alloyed glassy
powders implies the formation of the same glass state.
 2002 Elsevier Science B.V. All rights reserved.

Keywords: Amorphous materials; Transition metal alloys; Mechanical alloying; Magnetic measurements

1. Introduction tions [23–25]. Since the pioneering investigation of Koch


et al. [3] in 1983, numerous numbers of amorphous alloys
For more than 5000 years, most metallic alloys have and metallic glasses were fabricated by the MA method
been fabricated by melting and casting techniques [1]. The [26–42]. Accordingly, the term MA is becoming increas-
mechanical alloying (MA) [2,3] process is a rather new ingly common in both metal science and glass science.
comer to the world of metallurgy and materials science. In the recent years, metallic glassy soft magnetic
Whether it can be called a revolution or simple evolution, materials (see, for example, Refs. [43–45]) have received
the MA method has been successfully employed for much attention due to their unique properties and their
producing wide varieties of advanced materials, including promising applications. These materials offer the oppor-
metal nitrides [4,5], carbides [6,7], hydrides [8,9], silicates tunity to decrease transformer core losses. In particular a
[10], equilibrium [11] and nonequilibrium [12] phases, large elastic flexibility guarantees excellent insensitivity
composite and nanocomposite materials [13–16], and with respect to plastic deformations and a small electrical
nanocrystalline materials [17,18]. Amongst these useful conductivity reduces the eddy-current losses. Co-based
materials, metallic glassy alloys, with their unique short- glassy / amorphous alloys, with their non-magnetorestric-
range atomic order, have attracted research by many MA tion, possess excellent soft magnetic properties that pro-
scientists [19–22] due to the desirable properties that make pose them as core and sensor materials for inductive
them pioneering materials for several industrial applica- applications. Binary Co–Ti glassy alloys, which do not
show any deep eutectic compositions in the equilibrium
phase diagram [46] have not received much attention
*Corresponding author. Present address: Faculty of Engineering, Al-
Azhur University, Nasr City 11371, Cairo, Egypt. because of the difficulties of preparation by the conven-
E-mail address: msherif@mst1.mist.com.eg (M. Sherif El-Eskan- tional melt-spinning method. The only report for the
darany). formation of an ordinary amorphous Co 78 Ti 22 alloy by the

0925-8388 / 02 / $ – see front matter  2002 Elsevier Science B.V. All rights reserved.
PII: S0925-8388( 02 )00929-5
M. Sherif El-Eskandarany et al. / Journal of Alloys and Compounds 350 (2003) 232–245 233

way of rapid solidification from the molten metallic alloy contamination contents in the final product (86 ks) of MA
is that of Inoue et al. in 1980 [47]. powders are 0.14–0.22 wt% and 0.34–0.60 wt%, respec-
In the present study, the glass-forming ability of me- tively. Annealing the multilayered composite MA powders
chanically alloyed Co 1002x Ti x has been investigated. The was performed at the desired temperatures in well-sealed
effect of annealing on the structure and stability of the DSC Al-cells under continuous flow (2.5 ml s 21 ) of
multilayered Co / Ti composite powders, which are ob- purified argon flow. As soon as the annealing procedure
tained during the early stage of ball milling, has been was achieved, the DSC was rapidly cooled and the samples
studied. For the purpose of the present work, X-ray were subsequently removed from the DSC heating
diffraction, scanning and transmission electron micro- chamber. The as-annealed samples were structurally ana-
scopy, magnetic measurements, differential scanning lyzed by means of XRD, TEM / EDS and VSM techniques.
calorimetry and differential thermal analysis have been SEM / EDS technique was employed to observe the metal-
employed to monitor the effect of MA time on the lographical and compositional changes of the as-annealed
structural changes and the thermal stabilities of the fabri- MA particles.
cated glassy alloy powders.

3. Results

2. Experimental procedure 3.1. Structural changes with the MA time

Pure elemental powders (99.9% or better) of Co (25 The XRD patterns of mechanically alloyed Co 1002x Ti x
mm) and Ti (50 mm) were used as the starting reactant powders are shown in Fig. 1 after 86 ks of MA time (final
materials of MA. The powders were firstly balanced to product). Clearly visible haloes are obtained, especially in
give the nominal composition of Co 1002x Ti x (x525, 31, the Co-rich side (x,50 at%), indicating the completion of
33, 50 and 67 at%), mixed, charged into tempered chrome
steel vials (250 ml in volume) and then 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 MA experiments
were performed in a planetary ball mill (Fritsch P5) at a
rotation speed of 2.1 s 21 . In order to avoid an increase in
the vial temperature, the milling procedure was period-
ically interrupted every 1.8 ks and then halted for 3.6 ks
under a continuous flow of air. The later process was
necessary to eliminate the buildup of unprocessed powder
deposits at dead or abandoned spots inside the vials. The
ball milling was stopped after selected milling times and a
small amount (about 500 mg) of the MA samples were
taken from the vials in the glove box. The as-milled
samples were characterized by means of X-ray diffraction
(XRD) employing Cu Ka radiation, transmission electron
microscopy (TEM), using a 300-kV field emission micro-
scope, scanning electron microscopy (SEM) using 15-kV
field emission electron microscope, differential scanning
calorimetry (DSC) and differential thermal analysis (DTA)
at a constant heating rate of 0.67 K s 21 and under flow of a
purified argon gas (99.999 wt%). The magnetic polariza-
tion (Js ) of the ball-milled powders was measured at room
temperature, using a vibrating sample magnetometer
(VSM) with a maximum applied magnetic field of 670
kA m 21 . The coercive force was measured with a B–H
loop tracer. Energy dispersive spectroscopy (EDS), using
an electron beam of |5 nm has been used for analyzing
the concentration of the alloying elements, and the degree
of Fe contamination in the milled powders. In addition, the
oxygen content was determined by the helium carrier Fig. 1. XRD patterns of mechanically alloyed Co 1002x Ti x powders after
fusion–thermal conductivity method. The iron and oxygen 86 ks of the ball milling time.
234 M. Sherif El-Eskandarany et al. / Journal of Alloys and Compounds 350 (2003) 232–245

the solid-state reaction and the formation of amorphous phase maintains its unique structure and does not transform
alloys (Fig. 1a–c). The rather irregularity for the position into any other metastable / stable phase even after longer
of the first halo maximum may be attributed to the surface MA time (86 ks), as displayed in Fig. 2d.
oxidation of the amorphous powders that occurred during The bright field images (BFIs) of Co 75 Ti 25 powders at
handling the powders outside the glove-box. For x550 and the early stage of milling are shown in Fig. 3. After a few
67, the end-products are mixtures of an amorphous phase kiloseconds of MA time (1.8 ks), the metallic Co and Ti
coexisting with a small volume fraction of unprocessed powders tend to agglomerate and form large-grained
h.c.p.-Ti nanocrystalline powders (Fig. 1d and e). It can composite powders (Fig. 3a). These agglomerated particles
then be concluded that the mechanically alloyed Co–Ti are continuously subjected to impact and shear forces upon
binary system exhibits a considerably wide amorphization increasing the MA time. The existence of numerous faults
range, extending from 25 to 67 at%. with grain boundary fringes and heavy dislocations in the
In order to understand the amorphization process and the boundary is the general feature for the sample, which was
structural changes that take place upon ball milling the ball-milled for 7.2 ks (Fig. 3b).
alloying elements of Co and Ti powders, some samples The image of a high-resolution transmission electron
were taken after different MA times. The XRD patterns of microscope (HRTEM) of mechanically alloyed Co 75 Ti 25
mechanically alloyed Co 75 Ti 25 (selected here as a typical powders that was ball milled for 86 ks is shown together
example) powders are shown in Fig. 2 after selected MA with the corresponding selected area diffraction pattern
times. In contrast to the powders at the early stage of MA (SADP) in Fig. 4. The image shows maze pattern contrast
(0–7.2 ks), which are a mixture of polycrystalline h.c.p.- (Fig. 4a) with a typical halo-diffraction pattern of an
Co and h.c.p.-Ti (Fig. 1a,b), a broad diffuse and smooth amorphous phase (Fig. 4b).
halo appears after 32 ks of MA time (Fig. 2c), indicating In order to assess the distribution of the alloying
the formation of an amorphous phase. This amorphous elements (Co and Ti) in the powders of the final product
(86 ks), the local composition and the degree of homo-
geneity of the ball-milled sample have been examined by
the TEM / EDS technique. Some selected examined regions
for this sample are indexed in Fig. 4a and the corre-
sponding EDS analyses are listed in Table 1. Obviously,
the compositions do not fluctuate drastically from region to
region and the constituent elements of Co and Ti are
uniformly distributed in the powders, being rather close to
the starting nominal composition of Co 75 Ti 25 . The same
information could be obtained for the other compositions
of mechanically alloyed Co 1002x Ti x amorphous alloy
powders.

3.2. Metallographic changes with the MA time

Detailed SEM examinations were performed to under-


stand the metallography of mechanically alloyed
Co 1002x Ti x powders after different stages of MA. Fig. 5
shows the micrographs of the cross-sectional views of
Co 75 Ti 25 powders after selected MA times. After a few
kiloseconds of milling (1.8 ks), the starting ductile pow-
ders tend to agglomerate and form composite particles with
a thick multilayered structure (Fig. 5a). However, after 7.2
ks of MA time, the powders manifest a much finer
‘lamellar’ structure of the two constituents (Fig. 5b). In
addition, the layers are arranged in one orientation direc-
tion. Further milling (11 ks) enhances the shear and impact
forces generated by the balls and this leads to drastic
reduction of the layer thickness (Fig. 5c) so as to be hardly
visible after 86 ks of the MA time (Fig. 5e). The
disappearance of the layered-structure morphology implies
Fig. 2. XRD patterns of mechanically alloyed Co 75 Ti 25 powders after (a) the completion of the solid-state reaction and the formation
0, (b) 7.2, (c) 32, and (d) 86 ks of ball milling time. of a single phase.
M. Sherif El-Eskandarany et al. / Journal of Alloys and Compounds 350 (2003) 232–245 235

Fig. 3. BFI of mechanically alloyed Co 75 Ti 25 powders after (a) 1.8 and (b) 7.2 ks of ball milling time. The SADP that corresponds to the BFI in (b) is
presented in Fig. 3c.
236 M. Sherif El-Eskandarany et al. / Journal of Alloys and Compounds 350 (2003) 232–245

Fig. 4. (a) HRTEM and (b) the corresponding SADP of the final product (86 ks) of mechanically alloyed Co 75 Ti 25 powders. The EDS analyses of some
selected regions are listed in Table 1.

3.3. Magnetic measurements 3.4. Thermal stability

The typical hysteresis loops of the end-product (86 ks of The DSC scans of mechanically alloyed Co 1002x Ti x
MA time) of mechanically alloyed Co 1002x Ti x alloy pow- powders that were ball-milled for 86 ks (end-product) are
ders are shown in Fig. 6a. In spite of the samples of x550 displayed in Fig. 7. For all alloys in the composition range
and 67, the amorphous alloy powders in the Co-rich side (Fig. 7a–c) two separate reactions are taking place. The
(x,50) exhibit typical soft magnetic-type hysteresis loops. first reactions, which are endothermic, are due to the glass
Fig. 6b summarizes the magnetic properties of me- transition of the formed glassy phase, whereas the second
chanically alloyed Co 1002x Ti x powders, indexed by Js of sharp exothermic reaction is attributed to the crystallization
the end-products with respect to the Ti content, x. The Js of the glassy phase. It is worth mentioning that the
values of amorphous Co 75 Ti 25 , Co 69 Ti 31 and Co 67 Ti 33 appearance of the endothermic reactions during the DSC
alloy powders are 0.64, 0.61, and 0.42 T, respectively. scans, indicates the formation of glassy phase. In addition,
Contrary to these rather high values, the Js values of the the crystallization of the glassy phase takes place through a
amorphous Co 50 Ti 50 and Co 33 Ti 67 powders are 0.06 and single sharp exothermic reaction (Fig. 7a–c), suggesting
0.02 T, respectively. Obviously, Js is roughly decreasing that the formed glassy phase is single in structure and
linearly with increasing Ti content, x. homogeneous in composition. We will define the onset
temperature of the first and second reactions as glass
transition, T g and crystallization, T x temperatures, respec-
tively. In addition, the supercooled liquid region before
Table 1 crystallization, will be denoted as DT x (DT x 5 T x 2 T g ).
Local compositional EDS analysis of mechanically alloyed Co 75 Ti 25 Contrary to these glassy powders, the DSC scan for
powders that were obtained after 86 ks of milling time Co 33 Ti 67 (Fig. 7d) reveals a broad exothermic reaction,
Region Co (at%) Ti (at%) which lies at a relatively lower temperature (|727 K),
suggesting that the obtained amorphous phase is rather
1 75.7 24.3
2 76.3 23.7 heterogeneous. The absence of T g implies that the formed
3 75.2 24.8 metastable phase is an ordinary amorphous alloy.
4 74.7 25.3 In order to get more information about the progress of
5 75.8 24.2 the solid-state reaction during the several stages of MA,
The selected analytical zones are indexed in Fig. 4 (see the main text). selected ball-milled Co 1002x Ti x powders at the early,
M. Sherif El-Eskandarany et al. / Journal of Alloys and Compounds 350 (2003) 232–245 237

Fig. 5. SEM micrographs of the cross-sectional view for mechanically alloyed Co 75 Ti 25 powders after ball milling for (a) 1.8, (b) 7.2, (c) 11, and (d) 86 ks.

intermediate and final stages of MA were thermally Surprisingly, the endothermic reaction, which is clearly
analyzed. The DSC scans of the ball-milled powders, visible in the DSC scan of Fig. 8, disappears during the
exemplified by Co 75 Ti 25 powders are presented in Figs. 8 intermediate stage of the MA processing time (Fig. 9a,b).
and 9 after different MA times. Fig. 8 displays the DSC Moreover, the second exothermic reaction, which takes
scan of Co 75 Ti 25 powders after 11 ks of MA time. The place sharply in the DSC run of the 11 ks sample (Fig. 8),
scan reveals three separate reactions that take place: at 661 tends to be broadened with remarkable shift to the high
K (peak temperature of the first broad exothermic re- temperature side (857 K) upon MA for 22 ks (Fig. 9a).
action), at 765 K (onset temperature of the endothermic This crystallization reaction, however, becomes rather
reaction) and at 833 K (onset temperature of the second pronounced after 43 ks of MA time (Fig. 9b) and main-
sharp exothermic reaction). We should emphasise that the tains its tendency for shifting to the higher crystallization
first exothermic peak is a temporary reaction and com- temperature (T x ) side (876 K). The endothermic reaction
pletely disappeared during the second heating run of the corresponding to the glass transition temperature, T g ,
sample, which was previously annealed at 710 K. Contrary appears at 833 K after 65 ks of MA time (Fig. 9c) and
to this low temperature reaction, the last two peaks becomes more visible upon ball milling for 86 ks, when
permanently remain during the second DSC scan. This low the crystallization reaction becomes sharp (Fig. 9d). It is
temperature peak does not appear anymore in the DSC worth mentioning that the DT x of the end-product (86 ks)
scans of the samples at the intermediate (22–43 ks) and exhibits an extraordinary large value of 57 K for a binary
final (65–86 ks) stages of MA, as shown in Fig. 9. metallic glassy alloy. The T g was confirmed by heating the
238 M. Sherif El-Eskandarany et al. / Journal of Alloys and Compounds 350 (2003) 232–245

Fig. 7. DSC curves of the end-product (86 ks) of as-mechanically alloyed


Co 1002x Ti x powders for (a) Co 75 Ti 25 , (b) Co 67 Ti 33 , (c) Co 50 Ti 50 , and (d)
Co 33 Ti 67 glassy alloys.

of the existence of a glassy phase (Fig. 10a). In addition,


its hysteresis B–H loop (Fig. 11a) shows rather high
values of Js and coercive Hc , being very close to those
values for the powders at the early stage of MA, implying
the absence of any glassy phase.
Fig. 6. (a) Hysteresis loops of as-mechanically alloyed Co 1002x Ti x The XRD pattern of the sample that was taken after the
powders of (i) Co 75 Ti 25 , (ii) Co 69 Ti 31 , (iii) Co 67 Ti 33 , (iv) Co 50 Ti 50 , and completion of the first exothermic reaction (Fig. 8) reveals
(v) Co 33 Ti 67 amorphous alloys. (b) Magnetic polarization of the end- a typical halo-diffuse pattern of an amorphous phase (Fig.
product of MA as a function of Ti concentration (x). 10b). In addition, the hysteresis B–H loop of this sample
(Fig. 11b) shows lower values (in comparison with sample
samples to a temperature just above T g and then cooling [I) of Js (0.68 T) and coercive Hc (3.13 kA m 21 ), being
down to about 320 K. Then, the second and third heating slightly higher than those of the glassy Co 75 Ti 25 sample
runs were performed to confirm the reproducibility of T g obtained after 86 ks of MA time (see Fig. 6a(i)). This may
and to establish a base line. The T g always appears at the be attributed to the existence of unreacted elemental fine
same temperature (832.8–833.2 K) for all the three heating particles of Co in the as-annealed sample. The local
runs. compositional analysis of this annealed sample has been
We shall go back to the scan in Fig. 8, with its puzzling examined by the TEM / EDS technique. Aside from the
low-temperature exothermic peak and thermal stability. In featureless fine structure (Fig. 12a) and its correlated halo-
order to explore and interpret the roots of the fully diffuse SADP (Fig. 12b) of an amorphous phase, the
separated exothermic reactions, three separate samples compositions of some selected regions (see the table in
(denoted as [I, [II and [III), which had been originally Fig. 12) have close values, being very close to the starting
ball-milled for 11 ks, were independently heated to the nominal composition.
desired temperatures for several investigations. The XRD The XRD pattern of the sample heated to 993 K (Fig.
pattern reveals sharp Bragg-peak reflections that corre- 10c) reveals an f.c.c. structure with a lattice parameter (a 0 )
spond to elemental Co and Ti crystals, with no indication of 0.36175 nm, in good agreement with the reported value
M. Sherif El-Eskandarany et al. / Journal of Alloys and Compounds 350 (2003) 232–245 239

Fig. 8. Typical DSC curve of as-mechanically alloyed Co 75 Ti 25 powders


for 11 ks of the ball milling time. Points [I, [II and [III refer to the
samples that are taken for SEP/ EDS, XRD, VSM and TEM investigations
(see Figs. 10–13). Whereas T TEGFR is the peak temperature of the Fig. 9. DSC thermograms of mechanically alloyed Co 75 Ti 25 powders
amorphization peak, T g and T x are the onset temperatures of glass after (a) 22, (b) 43, (c) 65, and (d) 86 ks of the ball milling time.
transition and crystallization temperatures, respectively. The shaded area
under the DSC scan displays the structural changes that take place upon
heating these multilayered composite powders in the DSC. The crystalline at 710 K for 1.8 ks, a solid-state diffusion reaction is
composite Co 75 Ti 25 powders, which cover a temperature range extending taking place and a glassy phase is yielded. We shall call
from room temperature up to 696 K, are transformed into solid–amor- this process a thermally enhanced glass formation reaction
phous upon annealing at temperature II (710 K) which melts (glass
transition) and transforms into liquid–amorphous in an area which
(TEGFR).
extended from 768 to 831 K. It then gradually transforms into the most The thermal stability of mechanically alloyed Co 75 Ti 25
stable phase of f.c.c.-Co 3 Ti upon heating through a single exothermic powders, indexed by the T g , and crystallization tempera-
reaction that extends from 981 K. ture, T x are presented as a function of MA time in Fig. 14a.
In the figure, the T TEGFR (peak temperature of the low
of the ordered phase of f.c.c.-Co 3 Ti [48]. Furthermore, the exothermic temperature reaction in Fig. 8), which takes
hysteresis B–H loop (Fig. 11c) does not imply any soft place during the early stage of milling, is independent of
magnetic properties, with a low value of Js (0.23 T). MA time and disappears after 11 ks of the MA time. At
In addition to the above analyses, the melting (T m ) and this early stage of milling, T g appears in a temperature
liquidus (T l ) temperatures of the as-annealed sample at 710 range of 763–766 K and surprisingly disappeared as soon
K have been examined, using the DTA technique. The as T TEGFR disappears. The T x of the glassy phase obtained
typical DTA curve of the annealed sample is placed in Fig. at the early stage of milling slightly changes with increas-
13a together with the DTA scan of the glassy powders that ing MA time. It, however, changes dramatically during the
were obtained after 86 ks of MA time (Fig. 13b). The intermediate stage, indicating continuous compositional
curve a shows a single endothermic reaction which starts at changes of the mechanically alloyed powders. At the end
1425 K (T m ) and ends at 1480 K (T l ), in excellent of the final stage of milling (65–86 ks), T x tends to reach a
agreement with the reported values of f.c.c.-Co 3 Ti [46]. value of 889 K, implying the completion of the me-
These values are close to those of the final product (1431 chanically induced glass formation reaction (MIGFR) and
and 1476 K, respectively) (Fig. 13b). However, the T g /T l the formation of a single homogeneous glassy phase.
value of the as-annealed sample (0.52) is considerably Moreover, T g which disappeared during the intermediate
lower than that of the final product (0.56). Therefore, it and the mid final stages of milling, appears during the end
can be concluded that when pure multilayered Co / Ti of this final stage (76–86 ks). DT x of the glassy phase
composite particles are subjected to isothermal annealing obtained at the final stage of milling has a rather large
240 M. Sherif El-Eskandarany et al. / Journal of Alloys and Compounds 350 (2003) 232–245

Fig. 11. Hysteresis loops of previously mechanically alloyed Co 75 Ti 25


powders for 11 ks and then heated to (a) 473, (b) 710, and (d) 993 K (see
Fig. 6).

reaction during DSC measurements of the powders at the


Fig. 10. XRD patterns of previously mechanically alloyed Co 75 Ti 25
early stage of milling, which is previously shown in the
powders for 11 ks and then heated to (a) 473, (b) 710, and (d) 993 K (see
Fig. 6). DSC scan of x525 (see Fig. 8). DHTEGFR tends to decrease
monotonically during the first few kiloseconds of milling
and approaches minimum values, ranging from 20.39
value of 56 K. This value is considerably lower than that kJ mol 21 (x550) to 20.56 kJ mol 21 (x525). However, it
of the glassy phase, which is obtained by annealing the increases with increasing MA time and completely van-
powders of the early stage of milling (63 K). This may be ished towards the end of the MA time (Fig. 15a). In
attributed to the introduction of oxygen and iron contami- contrast to the DHTEGFR at the Co rich side (x#31),
nations into the powders (0.14 and 0.60 wt%, respectively) DHTEGFR at the Ti rich side (x$31), has lower values and
upon longer ball milling (86 ks). it takes longer MA time to disappear. It is worth mention-
The T g , T x , T m and T l of the end-product of me- ing that the low temperature peak due to the TEGFR does
chanically alloyed Co 1002x Ti x powders are presented in not appear during DSC scans of the composite powders at
Fig. 14b as a function of Ti content, x. The mechanically the Ti-rich side, which already have formed thick layers
alloyed glassy powders in the composition range 25$x# (.500 mm). Increasing the MA time does not improve the
50 exhibit large values of DT x (over 50 K), especially in metallographical characterizations, and the powders still
the Co-rich side (x#25). Both T g and T x depend on T m contain many thick layers even after longer MA time, as
and T l , which change markedly with changing Ti con- long as 86 ks. Accordingly, DHTEGFR for Co 33 Ti 67 (x567)
centrations. multilayered powders has zero value and does not change
Further information on the amorphization and crys- with increasing MA time, as displayed in Fig. 15a.
tallization reactions is given by the enthalpy change of During the early stage of MA (11 ks) the DHx slightly
glass formation (DHTEGFR ) via thermally enhanced glass changed with changing MA time (Fig. 15b). However, it
formation reaction TEGFR and enthalpy change of crys- drastically decreased to 22.19 kJ mol 21 (x525) to 20.68
tallization (DHx ) and presented as a function of MA time kJ / mol (x567) upon further ball milling time (11–43 ks),
in Fig. 15a and b, respectively. For all compositions indicating an increase in the volume fraction of the formed
(25#x#50) the values of DHTEGFR were directly mea- glassy phase. During the last stage of MA (65–86 ks) DHx
sured from the area under the low temperature exothermic tends to reach saturation values, ranging from 22.35
M. Sherif El-Eskandarany et al. / Journal of Alloys and Compounds 350 (2003) 232–245 241

Fig. 12. (a) HRTEM, and (b) the correlated SADP of previously mechanically alloyed Co 75 Ti 25 powders for 11 ks and then heated to 710 K. The
corresponding EDS analyses are listed in a table inside the figure.

kJ mol 21 (x525) to 20.96 kJ mol 21 (x50.67), indicating this system, which does not show any deep eutectic
the completion of the MA processing time. compositions [46]. The first attempt at preparing an
ordinary amorphous phase of Co 78 Ti 22 alloy was demon-
strated by Inoue et al. [47] when they reported the
4. Discussion possibility of forming an amorphous phase in a very
narrow range (between 21 and 23 at% Ti), using a melt-
In contrast to the Co–Zr binary system, with its several spinning method. The present study confirms that a solid-
deep eutectic compositions [49] which allows a wide state amorphization reaction takes place upon high-energy
glass-formation range (see, for example, Ref. [50]), the ball milling elemental powders of Co 1002x Tix under an
formation of a glassy phase in the Co–Ti binary system argon gas atmosphere at room temperature. Either a
has never, so far as we know, been reported in this system. thermally enhanced glass formation reaction (TEGFR) or a
This can be understood by looking at the phase diagram of mechanically induced glass formation reaction (MIGFR)
242 M. Sherif El-Eskandarany et al. / Journal of Alloys and Compounds 350 (2003) 232–245

Fig. 13. DTA curves of (a) as-mechanically alloyed Co 75 Ti 25 powders for


11 ks and then heated to 710 K, and (b) as-mechanically alloyed Co 75 Ti 25
powders for 86 ks. In the figure, T m is the melting temperature, whereas
T l is the liquidus temperature.

takes place when using the ball milling technique for


synthesizing glassy Co 1002x Ti x (25#x#50) powders.
When the elemental powders are mechanically alloyed for
a short MA time (1.8–3.6 ks), almost all the initial Co and
Ti powders are agglomerated to form large composite
particles, which contain large grains of the reactant materi-
als (Fig. 5a). Further milling time (3.6–11 ks) leads to
disintegration of these agglomerated powders and to
refinement of their ‘lamella’ so that the number of layers
per individual particle and their interfaces are monotonical-
ly increased. The continuous intensive impact and shear Fig. 14. (a) Dependence of the T TEGFR , T g and T x and DT x on the MA
forces that are generated by the milling tools (balls) create time during ball milling a mixture of Co 75 Ti 25 powders. (b) Correlation
clean well-developed multilayered Co / Ti composite par- between T g , T x , T m and T l , and Ti content of glassy Co 1002x Ti x alloy
powders that were obtained after 86 ks of the MA time. DT x TEGFR and
ticles typical of sputtered or evaporated diffusion couples
DT x MIGFR refer to the supercooled liquid region of the glassy phase that is
[51–53]. During the DSC experiments of these formed by way of thermally enhanced glass formation reaction (TEGFR)
multilayered powders, two clear separate exothermic peaks and mechanically induced glass formation reaction (MIGFR) processes,
appear (Fig. 8). The first exothermic reaction that appears respectively.
at 661 K takes place due to a solid-state reaction between
the thin elemental layers of Co and Ti and this leads to the
formation of a homogeneous (Fig. 12) glassy phase (Fig. also reported for some mechanically alloyed metallic
10b) with a composition close to the starting nominal powders (see, for example, Refs. [27,30,35]). Indeed, post-
composition (see the EDS analytical results in Fig. 12). annealing the composite powder particles enhances the
The formation of a glassy phase via post-annealing me- solid-state diffusion between the diffusion couples and this
chanically alloyed powders is not unique for the Co–Ti leads to a speeding up of the rate of diffusion at the clean
binary system. It has been recently shown by El-Eskan- Co / Ti boundaries. Thus, the free energy changes drastical-
darany and Inoue [54] that a glassy phase of Cu 33 Zr 67 ly from a non stable (the starting reactant materials) to a
alloy powders can be also obtained by isothermal-anneal- more stable phase (glass). For a successful TEGFR, the
ing the multilayered composite powders at the early and applied annealing temperature must be well below the
intermediate stage of milling. Such a phenomenon was temperatures that suppress the nucleation and growth of
M. Sherif El-Eskandarany et al. / Journal of Alloys and Compounds 350 (2003) 232–245 243

layers (which became very narrow) increases. Accordingly,


DHTEGFR decreases and approaches minimum values (Fig.
15a). It is worth mentioning that the powders at this early
stage of milling do not contain any mole fractions of the
glassy phase (Figs. 2b and 3a,b), indicating that the glassy
phase of this stage is obtained only by the TEGFR process.
During the next stage of milling (22–32 ks), a considerable
volume fraction of the glassy phase is formed due to
MIGFR (Fig. 1c) so that the glassy phase of this stage is
obtained by both TEGFR and MIGFR.
We denote the heat formation of a glassy phase via
MIGFR by the term DHMIGFR . However, DHMIGFR is not a
measurable value; it can be estimated during the final stage
of milling (where glassy formation peaks are absent, i.e.
DHMIGFR 50), using the following relations:

DHfor 5 DHG 1 DHx (1)

Here, is the reported value of the enthalpy of formation of


an intermetallic compound at specific concentrations of
Co 1002x Ti x , based on Miedema’s model [55], DHG is the
heat formation of a glassy phase and DHx is the enthalpy
of crystallization. During the final stage of milling, where
the glassy phases are formed due to MIGFR, Eq. (1) can
be written as:
DHfor 5 DHMIGFR 1 DHx MIGFR (2)

DHMIGFR is the total heat formation of Co 1002x Ti x glassy


phase via the MIGFR process (ball milling) and DHx MIGFR
is the enthalpy of crystallization for the glassy phases that
are formed by the MIGFR. In fact, this value is very
difficult to be estimated at the intermediate stage of milling
in which the glassy phase results from TEGFR and
MIGFR together.
Then, DHMIGFR can be given as follows:

DHMIGFR 5 DHfor 2 DHx MIGFR (3)

The measured values of DHTEGFR (heat of formation of


glassy alloys), the estimated values of DHMIGFR of glassy
Co 1002x Ti x (x525, 33 and 50) and the relative heat of
Fig. 15. (a) Dependence of DHTEGFR and (b) DHx on the MA time during glassy formation ratio (DHMEGFR /DHTEGFR ) are listed in
ball milling a mixture of Co 1002x Ti x powders.
Table 2. At all compositions, DHMEGFR /DHTEGFR , ratio has

the equilibrium phase, i.e. the crystallization temperature. Table 2


The remarkable decrease in the magnetic polarization of Heat formation of glassy Co 1002x Ti x alloy powders via thermally en-
hanced glass formation reaction, TEGFR (DHTEGFR ) and mechanically
as-annealed powders (Fig. 11b) suggests the completion of induced glass formation reaction, MIGFR (DHMIGFR ), and the relative
the solid-state reaction and the decreasing of Co-free atoms heat of glassy formation ratio
in the composite powders upon annealing at 710 K. DH MIGFR
Ti concentration DHTEGFR DHMIGFR ]]
In the present study, the term DHTEGFR refers to the heat (at%) (kJ mol 21 ) (kJ mol 21 )
DH TEGFR

formation of the glassy phase (enthalpy of glassy forma-


25 20.56 227.65 49.4
tion) via annealing the multilayered milled powders. This 31 20.44 232.77 74.5
value is measured directly from the area under the amor- 33 20.43 235.82 83.3
phization peak during the DSC scans of the ball-milled 50 20.39 240.53 103.9
alloy powders at the early and intermediate stages of MA. The values of DHTEGFR were directly measured from the DSC scans of
Once the thickness of Co and Ti layers decreases with the mechanically alloyed powders, whereas DHMIGFR were estimated from
increasing the MA time (Fig. 5), the number of these fresh Eq. (2) (see the main text).
244 M. Sherif El-Eskandarany et al. / Journal of Alloys and Compounds 350 (2003) 232–245

large values, suggesting that the glassy phases are mainly the elemental mixed powders are milled for longer MA
formed by MIGFR. time (86 ks), the glass formation reaction is taking place
In the MIGFR process, the solid-state reaction takes due to mechanically induced solid state reaction. The final
place in the same manner as occurs in the TEGFR process, products of the Co-rich glassy alloys, that were obtained
except with a lower diffusion rate and a longer MA time after ball milling for 86 ks, exhibit good soft magnetic
(Fig. 2). This is attributed to the milling temperature that is properties. The fabricated glassy alloys have large values
assumed to be far below than in TEGFR (710 K). At the for the supercooled liquid region before crystallization,
initial milling stage, crystalline elemental powders are DT x (larger than 50 K).
mechanically crushed and fresh surfaces are rapidly
created. Kneading of such ground powders enhances the
atomic diffusion of elemental atoms and step-by-step local References
alloying [42]. Thus, the interdiffusion between Co and Ti
layers occurs slowly. Since MA introduces many vac- [1] M. Sherif El-Eskandarany, Mechanical Alloying for Fabrication of
ancies, lattice defects, grain boundaries and surfaces, the Advanced Engineering Materials, 1st Edition, William Andrew, New
ball-milled powders store a large amount of mechanical- York, 2001, p. 5.
strain energy [36,56,57]. Increasing the MA time during [2] J.S. Benjamin, Metall. Trans. A 1 (1970) 2943.
the intermediate stage of milling leads to a continuous [3] C.C. Koch, O.B. Cavin, C.G. McKamey, J.O. Scarbrough, Appl.
Phys. Lett. 43 (1983) 1017.
movement of the milling media which leads to an excess [4] M. Sherif El-Eskandarany, M. Omori, T.J. Konno, K. Sumiyama, T.
of the lattice imperfections that assist the diffusion be- Hirai, K. Suzuki, Metall. Trans. A 29 (1998) 1973.
tween the Co–Ti diffusion couple which have already [5] D. Wexler, A. Calka, A.Y. Mosbah, J. Alloys Comp. 309 (2000) 201.
negative heats of mixing [55]. In the milling process, [6] M. Sherif El-Eskandarany, Metall. Trans. A 27 (1996) 2374.
friction between balls, balls and the lining surface of a vial [7] M. Sherif El-Eskandarany, J. Alloys Comp. 305 (2000) 219.
[8] M. Sherif El-Eskandarany, H.A. Ahmed, K. Sumiyama, K. Suzuki,
generate frictional heat. As a result of the ball motion and
J. Alloys Comp. 218 (1995) 36.
their collisions, the interdiffusion in the multilayered [9] I.G. Konstanchuk, E.Y. Ivanov, B.B. Bokhonov, V.V. Boldyrev, J.
powders is enhanced. In fact, such heat not only acceler- Alloys Comp. 319 (2001) 290.
ates diffusion of the constitutional atoms of the powder [10] W. Lee, J. Lee, J.D. Bae, C.S. Byun, D.K. Kim, Scripta Mater. 77
mixture, but also contributes to thermal annealing of the (2001) 97.
[11] M. Krasnowski, H. Matyja, J. Alloys Comp. 319 (2001) 296.
unprocessed multilayered composite powders, which may
[12] A.M. Korsunsky, A.I. Salimon, I. Pape, A.M. Polyakov, A.N. Fitch,
remain in the mixture. J. Alloys Comp. 315 (2001) 217.
In contrast to the crystallization process of melt-spun [13] Y.D. Kim, S.-T. Oh, K.H. Min, H. Jeon, I.-H. Moon, Scripta Mater.
Co 78 Ti 22 which takes place through two steps [46], the as 44 (2001) 293.
MA–Co 1002x Ti x glassy alloys transform into the equilib- [14] Z.Y. Liu, N.H. Loh, K.A. Khor, S.B. Tor, J. Alloys Comp. 311
(2001) 13.
rium phases through single sharp exothermic reactions.
[15] M. Sherif El-Eskandarany, J. Alloys Comp. 279 (1998) 263.
This is attributed to the high capability of the mechanically [16] M. Sherif El-Eskandarany, M. Omori, T.J. Konno, K. Sumiyama, T.
induced solid state reaction to homogenize the powders at Hirai, K. Suzuki, Metall. Trans. A 32 (2001) 157.
an atomic scale during the different stages and to over- [17] K.W. Liu, F. Mucklich, Acta Mater. 49 (2001) 395.
come the precipitation of any short or medium range order [18] F. Wu, P. Bellon, A.J. Melmed, T.A. Lusby, Acta Mater. 49 (2001)
453.
in the glassy matrix. The formed glassy phase at the last
[19] A. Inoue, K. Matsuki, T. Masumoto, Mater. Trans. J. Inst. Metall. 31
stage of milling is homogeneous, indicated by the near (1990) 148.
values of Js , T g , T x and DHx of the powders. In addition, ¨
[20] J. Eckert, M. Seidel, N. Schlorke, A. Kubler, L. Schultz, Mater. Sci.
the EDX analyses for the final product of the glassy Forum 235–238 (1997) 23.
powders have not shown any significant concentration [21] M. Seidel, J. Eckert, L. Schultz, Mater. Sci. Forum 235–238 (1997)
29.
gradients or compositional fluctuation, indicating that the
[22] M. Sherif El-Eskandarany, A. Inoue, Metall. Trans. A 33 (2002)
obtained glassy phases are uniform and homogeneous at 135.
the atomic scale. ¨
[23] A. Inoue, in: M. Magini, F.H. Wohlbier (Eds.), Bulk Amorphous
Alloys: Practical Characteristics and Applications, 1st Edition, Trans
Tech Publications, Switzerland, 1999, pp. 140–141.
5. Conclusions [24] J. Gottschall, Mater. Trans. J. Inst. Metall. 42 (2001) 548.
[25] A. Inoue, in: A. Inoue, K. Hashimoto (Eds.), Amorphous and
Nanocrystalline Materials: Preparation, Properties and Applications,
Glassy Co 1002x Ti x alloy powders have been synthesized 1st Edition, Springer, Berlin, 2001, pp. 47–48.
by high-energy ball milling the elemental powders at room [26] R.B. Schwarz, C.C. Koch, Appl. Phys. Lett. 49 (1986) 146.
temperature, using the mechanical alloying method. During [27] R.B. Schwarz, R.R. Petrich, J. Less-Common Metals 140 (1988)
171.
the early and intermediate stages of milling, post-annealing
[28] M. Sherif El-Eskandarany, K. Aoki, K. Suzuki, J. Less-Common
of the mechanically deformed Co / Ti multilayered compo- Metals 167 (1990) 113.
site powders at 710 K leads to the formation of a glassy [29] M. Sherif El-Eskandarany, K. Aoki, K. Suzuki, Appl. Phys. Lett. 60
phase (thermally enhanced glass formation reaction). When (1992) 1562.
M. Sherif El-Eskandarany et al. / Journal of Alloys and Compounds 350 (2003) 232–245 245

[30] M. Sherif El-Eskandarany, K. Aoki, K. Suzuki, J. Appl. Phys. 71 [43] T. Mizushima, A. Makino, A. Inoue, J. Appl. Phys. 83 (1998) 6329.
(1992) 2924. [44] A. Inoue, B. Shen, Mater. Trans. J. Inst. Metall. 43 (2002) 1230.
[31] M. Sherif El-Eskandarany, K. Aoki, K. Suzuki, J. Appl. Phys. 72 [45] H. Koshiba, A. Inoue, Mater. Trans. J. Inst. Metall. 42 (2001) 2572.
(1992) 2665. [46] T.B. Massalski (Ed.), 2nd Edition, Binary Alloy Phase Diagrams,
[32] M. Sherif El-Eskandarany, Metall. Trans. A 27 (1996) 3267. Vol. 2, ASM International, Washington, DC, 2000, p. 1251.
[33] M. Sherif El-Eskandarany, K. Aoki, K. Sumiyama, K. Suzuki, Appl. [47] A. Inoue, K. Kobayashi, C. Suryanarayana, T. Masumoto, Scripta
Phys. Lett. 70 (1997) 1679. Metall. 14 (1980) 119.
[34] M. Sherif El-Eskandarany, K. Sumiyama, K. Suzuki, Acta Mater. 45 [48] JADE card [23-0938.
(1997) 1175. [49] T.B. Massalski (Ed.), 2nd Edition, Binary Alloy Phase Diagrams,
[35] M. Sherif El-Eskandarany, J. Alloys Comp. 284 (1999) 295. Vol. 2, ASM International, Washington, DC, 2000, p. 1265.
[36] M. Sherif El-Eskandarany, K. Sumiyama, K. Suzuki, Acta Mater. 50 [50] K.H.J. Buschow, J. Less-Common Metals 85 (1982) 221.
(2002) 1113. [51] R.B. Schwarz, W.L. Johnson, Phys. Rev. Lett. 51 (1983) 415.
[37] M. Sherif El-Eskandarany, A. Inoue, J. Non-Cryst. Solids (2002) in [52] B.M. Clements, W.L. Johnson, R.B. Schwarz, J. Non-Cryst. Solids
press. 61–62 (1984) 817.
[38] M. Sherif El-Eskandarany, A. Inoue, Metall. Trans. A (2002) in [53] E.J. Cotts, W.J. Meng, W.L. Johnson, Phys. Rev. Lett. 57 (1986)
press. 2295.
[39] M. Sherif El-Eskandarany, A. Inoue, J. Non-Cryst. Solids (2002) in [54] M. Sherif El-Eskandarany, A. Inoue, Mater. Trans. J. Inst. Metall.
press. 43 (2002) 608.
[40] M. Sherif El-Eskandarany, W. Zhang, A. Inoue, Mater. Trans. J. Inst. [55] F.R. de Boer, R. Boom, W.C.M. Mattens, A. Miedema, A.K.
Metall. 43 (2002) 1422. Niessen, Cohesion in Metals–Transition Metal Alloys, North-Hol-
[41] M. Sherif El-Eskandarany, W. Zhang, A. Inoue, J. Mater. Res. land, Amsterdam, 1988, pp. 266.
(2002) in press. [56] A.E. Ermakov, E.E. Yurchikov, V.A. Barinov, Phys. Met. Metall. 52
[42] M. Sherif El-Eskandarany, J. Saida, A. Inoue, Acta Mater. 50 (1981) 50.
(2002) 2725. [57] A.W. Weeber, H. Bakker, F.R. Boer, Europhys. Lett. 2 (1986) 445.

Anda mungkin juga menyukai