Anda di halaman 1dari 6

Journal of Non-Crystalline Solids 404 (2014) 712

Contents lists available at ScienceDirect

Journal of Non-Crystalline Solids


journal homepage: www.elsevier.com/ locate/ jnoncrysol

Crystallization kinetics of Cu38Zr46Ag8Al8 bulk metallic glass in different


heating conditions
J. Cui a, J.S. Li a, J. Wang a,, H.C. Kou a, J.C. Qiao b, S. Gravier c, J.J. Blandin c
a
b
c

State Key Laboratory of Solidication Processing, Northwestern Polytechnical University, Xi'an 710072, Shaanxi Province, PR China
Universit de Lyon, CNRS, INSA-Lyon, MATEIS UMR5510, F-69621 Villeurbanne, France
SIMAP, INP Grenoble/CNRS/UJF, 38402, Saint-Martin d'Hres, Cedex, France

a r t i c l e

i n f o

Article history:
Received 16 April 2014
Received in revised form 17 July 2014
Available online 10 August 2014
Keywords:
Bulk metallic glass;
Crystallization kinetics;
Activation energy;
Kissinger and JohnsonMehlAvrami;
Local Avrami exponent

a b s t r a c t
The crystallization kinetics of Cu38Zr46Ag8Al8 bulk metallic glass in non-isothermal mode and isothermal mode
are investigated by differential scanning calorimetry. In non-isothermal conditions, the average value of
activation energy is determined by Kissinger equation, and the value is around 310 kJ/mol. The crystallization
enthalpy is about 28.69 J/g. In addition, the local Avrami exponent is adopted to describe the crystallization
process. In isothermal route, the average value of activation energy for crystallization is calculated by the Arrhenius
equation, and the value is about 451 kJ/mol. The crystallization enthalpy is about 1.63 J/g. And the Avrami
exponent n ranges from 4.10 to 4.74, which indicates that the crystallization mechanism is mainly governed by
constant nucleation rate. The created phases in the two conditions are different which can be conrmed by the
X-ray diffraction test, this result is in accordance with the different crystallization enthalpies, but it's different
with other investigations.
2014 Elsevier B.V. All rights reserved.

1. Introduction
Compared with their crystalline counterparts, bulk metallic glasses
(BMGs) have a series of special combination of structural and functional
properties, such as higher strength, larger elasticity, lower elastic
modulus and excellent corrosion resistance due to the lack of translational
or orientational long-range order [1,2]. Namely, there are no conventional
defects in bulk metallic glasses [38], such as grain boundary and
dislocations. Therefore, bulk metallic glasses are potential materials
which can be applied on structural and functional materials.
In recent years, Cu-based bulk metallic glasses have been widely
investigated due to their excellent glass forming ability, good thermal
stability, wide supercooled liquid region (SLR) and high mechanical
properties at room temperature, as well as the containing of relatively
low cost elements [911]. Therefore, it is an ideal candidate which can
be applied in near shape fabri cation of precise and complex-shaped
components by means of the thermal plastic forming (TPF) method.
During the process of TPF, crystallization of BMGs must be avoided,
since crystallization or partial crystallization of BMGs will lead to a
drastic degradation of mechanical properties at room-temperature
[12]. In other words, crystallization makes BMG parts unsafe. Based on
this discussion, if we want to avoid crystallization effectively, it's
meaningful to understand the crystallization kinetics of BMG.
Corresponding author. Tel.: +86 29 8846 0568; fax: +86 29 8846 0294.
E-mail address: nwpuwj@nwpu.edu.cn (J. Wang).

http://dx.doi.org/10.1016/j.jnoncrysol.2014.07.029
0022-3093/ 2014 Elsevier B.V. All rights reserved.

However, BMGs tend to transfer from amorphous state to crystalline


state by thermal annealing. Moreover, thermal stability always links to
the mechanical properties in BMGs. Thus, it is very necessary to
investigate the thermal stability and crystallization behavior in bulk
metallic glasses. In general, differential scanning calorimetry (DSC) or
differential thermal analysis (DTA) is widely used to research thermal
stability and crystallization behavior in amorphous materials [1318].
The properties for amorphous materials, such as glass transition
temperature (Tg), temperature corresponding to the onset of crystallization (Tx), crystallization peak temperature (Tp) and crystallization
enthalpy could be obtained from the DSC curves. Furthermore, apparent
activation energy could be calculated by relevant models from the DSC
data. In effect, thermal stability and crystallization behavior restrict the
applications of bulk metallic glasses.
It is well-known that there are two steps leading crystallization
during the process of TPF: continuous heating from low temperature
(non-isothermal mode) and annealing at a given temperature (isothermal mode). For these two modes of crystallization, all we are concerned
about are the activation energy of crystallization and the description of
whole crystallization process. There are different methods that can
be employed to describe activation energy of crystallization, such
as Kissinger method [19]and Friedman method [20] and many subject
matters are modeled to analyze the thermoanalytical kinetic patterns
[21,22]. It should be noted that JohnsonMehlAvrami (JMA) model is
always used to describe the crystallization process and gives good
results [2327], for example, Xie et al. [28] adopted this method in

J. Cui et al. / Journal of Non-Crystalline Solids 404 (2014) 712

Cu50Zr45Ti5 BMG. Moreover, this method also can be used for any kind
of phase transformation.
Therefore, the main purpose of this work is to investigate crystallization transformation kinetics in the non-isothermal and isothermal
modes of Cu 38 Zr46 Ag 8Al8 bulk metallic glass by DSC. In addition,
the research can provide some new insight to further understand
crystallization transformation kinetics, relevant thermal stability
and physical properties in bulk metallic glasses.
2. Experimental procedure
2.1. Sample preparation
The pre-alloyed ingot of Cu38Zr46Ag8Al8 (at%) BMG was prepared
by arc melting under a puried argon atmosphere in a water-cooled
copper crucible and in situ suction casting into a water-cooled
copper mold, then the rods of BMG with a diameter around 3 mm
were fabricated. Purity of the elements was above 99.9%. The ingot
was re-melted several times in order to ensure the homogeneity.
The DSC samples were mechanically cut from as-cast rods, and the
surfaces of samples were carefully polished and nally washed in
ethanol by an ultrasonic cleaning machine in order to remove the
surface oxidation before the DSC experiments. The amorphicity of
Cu38 Zr46 Ag8 Al 8 BMG was detected by means of X-ray diffraction
(XRD, DX-2700).
2.2. DSC experiments
The information of thermal properties and phase transformation
were determined by differential scanning calorimetry (DSC, Perkin
Elmer, Diamond) under a high purity owing argon gas atmosphere
with the owing rate of 20 ml/min and alumina pans were used as
sample holders, and the temperature accuracy is 0.01 K.
The non-isothermal DSC experiments were carried out by using
different steady heating rates from 2.5 K/min to 20 K/min. The isothermal
crystallization experiments were performed in the supercooled liquid
region (SLR) of Cu38Zr46Ag8Al8 BMG. The samples were heated up to
the annealing temperature at a heating rate of 20 K/min, then holding
at the selected temperatures until the crystallization completed, after
that, samples were cooled down to room temperature.
3. Results and discussion
3.1. Non-isothermal crystallization behavior
The crystallization peaks of DSC curves in Cu38Zr46Ag8Al8 BMG
performed at different heating rates are shown in Fig. 1. The heating

Fig. 1. The crystallization part of DSC curves in Cu38Zr46Ag8Al8 bulk metallic glass with
various heating rates.

rates are chosen as 2.5 K/min, 5 K/min, 10 K/min and 20 K/min, respectively. As can be seen from Fig. 1, crystallization only induces a single
pronounced exothermic peak at each heating rate, and the position of
peak shifts to higher temperature with the increase of heating rate.
Therefore, the crystallization process in this alloy is simple, and the
crystallization kinetics can be analyzed directly. By contrast with the
single exothermic peak, there are also complex exothermic peaks that
exist in other kinds of BMGs. For example, double exothermic peaks
can be obtained in classical VIT1 BMG [29], and three exothermic
peaks can be obtained in Cu52.5Ti30Zr11.5Ni6 BMG [30], all of these
peaks correspond to the multiplestage crystallization modes. For these
alloys, the crystallization kinetics is difcult to analyze, since all the
contributions have to be discussed separately. The XRD patterns of the
crystallized BMG are shown in Fig. 2, from which we can see that
there are two created phases in non-isothermal condition: CuZr phase
and AlAg phase, and three created phases in isothermal condition:
CuZr phase, AlAg phase and AgZr phase. The different created phases
in the two heating conditions can also be conrmed by the different
crystallization enthalpy, which have been done in the following sections.
And the result is different with the former research by Wei, et al. [31].
If we want to analyze the crystallization kinetics during the whole
process, the activation energy of the BMG in different crystallization
volume fraction must be obtained. Before that, the crystallization
volume fraction for non-isothermal crystallization, , should be
obtained, which can be deduced as a function of temperature from
DSC curves by using the following equation [3235]:
Z
Z

T
T0
T
T0

dHc =dT dT

dHc =dT dT

A0
A

where T0 and T are the temperatures at which crystallization begins


and ends in amorphous materials, dHc/dT is the heat capacity in a
constant pressure, and A0 and A are the areas under the DSC curves.
The relationship between crystallization fraction and temperature at
different heating rates is shown in Fig. 3, all the curves exhibit typical
S-shaped type in different heating rates. Finally, the samples from
metastable state transfer to a more stable state.
The crystallization quantity in non-isothermal condition can be
identied by calculating the area of crystallization enthalpy by using
the following equation:
H

1
Rh

T
T0

W T dT

where H is crystallization enthalpy, W(T) is heat ow, and Rh is heating


rate. From this equation, the crystallization enthalpy can be calculated;

Fig. 2. XRD patterns of crystallized Cu38Zr46Ag8Al8 bulk metallic glass in different


conditions.

J. Cui et al. / Journal of Non-Crystalline Solids 404 (2014) 712

Fig. 3. Relationship between crystallization fraction and temperature at different heating


rates as a guide to the eyes.

Fig. 5. Relationship between ln[ln(1)] and 1000/T of Cu38Zr46Ag8Al8 bulk metallic


glass at different heating rates as a guide to the eyes.

the crystallization enthalpy is 25.92 J/g at a heating rate of 2.5 K/min,


27.24 J/g at 5 K/min, 28.8 J/g at 10 K/min and 32.79 J/g at 20 K/min.
The result indicates that the crystallization enthalpy is independent of
heating rate, and it almost maintains the same level, which also means
that the created phases are the same at these heating rates.

at heating rate of 2.5, 5, 10, and 20 K/min, respectively. It can be seen


that in the whole process, the Avrami exponent is different with many
distinct stages. So the local Avrami exponent n() should be introduced.
In order to calculate the local Avrami exponent for non-isothermal
crystallization, we adopt the equation as follows [39]:

3.1.1. Activation energy


The apparent activation energy, Ea, for glass transition or crystallization of amorphous materials under continuous heating conditions can
be determined by the Kissinger equation [3638]:

ln

Rh
T p2

Ea
C
RT p

where T p is the temperature corresponding to the maximum of


crystallization peak (in here, it indicates the maximum temperature
value of each curve in Fig. 1), and the approximate straight line could
be obtained by using Kissinger method to t the experimental values,
the result is shown in Fig. 4. From the slope of this line, we can calculate
the activation energy; the value is about 310 kJ/mol in non-isothermal
heating conditions.
3.1.2. Crystallization mechanism
For non-isothermal conditions, the Avrami exponent n can be
obtained by plotting ln[ln(1)] vs. 1/T, which can be tted by a
straight line with a slope nEc/R, then the value of Avrami exponent, n,
can be calculated. Fig. 5 shows the curves of ln[ln(1)] vs. 1000/T

Fig. 4. Kissinger plot of Cu38Zr46Ag8Al8 bulk metallic glass in non-isothermal condition.


The peak temperatures in 4 heating rates are: 758 K, 768 K, 778 K and 790 K. The t has
R2 value of 0.998. The solid line is tted by Eq. (3), and the activation energy is about
310 kJ/mol.

R ln ln 1 
:
Ea 1=T

Based on this equation, the local Avrami exponents at heating rates


from 2.5 K/min to 20 K/min are calculated, the results are shown in
Fig. 6. It can be seen that the Avrami exponents at different heating
rates display the same tendency and they vary with the increase of
crystallization fraction during the whole process. Initially, when is
between 20% and 80% the n() is larger than 4, which means that the
nucleation rate increases with time, and the crystallization quantity
increases rapidly. When the reaches 80%, the n() value is back to
34, it means that the crystallization progresses with a decreasing
crystallization rate, the main reason is due to the lack of solvend.
3.2. Isothermal crystallization behavior
Isothermal crystallization of the metallic glass is carried out in the
SLR, i.e. above the glass transition temperature (Tg) and below the
temperature corresponding to the onset of crystallization (Tx). The
samples are heated up to the annealing temperature with the heating
rate of 20 K/min and then held until the crystallization is completed.
Finally, samples are cooled down to room temperature.

Fig. 6. Relationship between local Avrami exponent n() and crystallization volume
fraction at different heating rates as a guide to the eyes.

10

J. Cui et al. / Journal of Non-Crystalline Solids 404 (2014) 712

In isothermal crystallization condition, the annealing temperatures


are selected in the SLR of Cu38Zr46Ag8Al8 BMG due to the TPF method
as always performed in the SLR. As shown in Fig. 7, we can obtain the
Tg = 695 K and Tx = 785 K, so the annealing temperatures are chosen
as 755 K, 761 K, 767 K, 773 K and 779 K. The crystallization peaks of
each annealing temperature in DSC curves of Cu38Zr46Ag8Al8 BMG are
shown in Fig. 8, the same with the plots in non-isothermal condition.
The DSC curves only exhibit a single pronounced exothermic peak
after passing a certain incubation period, , which decreases with the
increase of annealing temperature. This phenomenon can be ascribed
to the higher atomic mobility at higher annealing temperature, which
can cause concentration uctuation necessary for large-scale crystallization to set in [39]. The same phenomenon can be found in previous
reports in Zr-based BMGs [4042] and Cu-based BMGs [30,43].
The crystallization volume fraction for isothermal crystallization, ,
can be deduced as a function of annealing time from DSC curves by
equation which is given by [32]:
Z t
dHc =dt dt
A
t
Z t0
0
5

A
dHc =dtdt
t0

where t0 and t are the time when crystallization begins and ends in
amorphous materials, dHc/dt is heat ow, and A0, A are also the areas
under DSC curves. Fig. 9 exhibits the relationship between crystallization fraction and annealing time. The plots all display an S-shaped
trend, and the crystallization process becomes slow with the decrease
of annealing temperature.
Almost the similar with non-isothermal condition, the crystallization quantity in isothermal condition can be identied by calculating
the area of crystallization enthalpy using the following equation:
H

1
Rh

t
t0

W t dt:

The crystallization enthalpy can be calculated by using this equation,


the results are 1.68 J/g at annealing temperature of 755 K, 1.47 J/g at
761 K, 1.68 J/g at 767 K, 1.62 J/g at 773 K and 1.71 J/g at 779 K. The
crystallization enthalpy maintains similar value in the isothermal
condition; the created phases are also the same.
3.2.1. Activation energy
In isothermal crystallization condition, the Arrhenius equation is
often utilized to evaluate the activation energy, Ea[29]:

t t 0 exp

Ea
RT


7

Fig. 7. DSC curve of Cu38Zr46Ag8Al8 bulk metallic glass (heating rate: 20 K/min). Glass
transition temperature (Tg) and onset crystallization temperature (Tx) are dened.

Fig. 8. The crystallization part of DSC curves in Cu38Zr46Ag8Al8 bulk metallic glass at
different annealing temperatures.

where t0 is a constant andR is the gas constant. The plots of ln(t())


versus 1/T at different crystallization volume fractions (from 0.1 to
0.9) are shown in Fig. 10, and the approximate straight lines can be
obtained by tting the experimental values using Arrhenius equation.
The activation energy can be obtained from the slope of tting curves
as shown in Fig. 11, the average value is about 451 23 kJ/mol in
isothermal heating condition. However, from the results we can see
that the activation energy decreases with the increase of crystallization
fraction. This tendency is due to the energy that consists of two components: one term corresponding to the nucleation and the other to the
growth. When crystallization progresses, the energy required for
nucleation can progressively disappear and then Ea decreases. Compared
with isothermal annealing condition, the sample in continuous heating
condition will be crystallized at a relative higher temperature, which
means that crystallization transformation from metastable state to
crystalline phases is easier at higher heating temperature.
3.2.2. Crystallization mechanism
The isothermal annealing time and the crystallization volume
fraction can be modeled by the JohnsonMehlAvrami (JMA) equation
to describe the transformation kinetics as follows [41,4446]:

n
t 1 exp K t

where is incubation time for nucleation. Taking the double logarithm


of Eq. (8), we can deduce the following equation [40,44]:
ln ln 1 t  n lnK n ln t :

Fig. 9. Relationship between crystallization volume fraction and annealing time at different annealing temperatures in Cu38Zr46Ag8Al8 bulk metallic glass as a guide to the eyes.

J. Cui et al. / Journal of Non-Crystalline Solids 404 (2014) 712

11

Fig. 10. Plots of Cu38Zr46Ag8Al8 bulk metallic glass for activation energy in isothermal
conditions. The ts have R2 values all above 0.99.

The JMA plots of ln[ ln(1 )] versus ln(t ) are shown in


Fig. 12 from which the experimental points can be tted to give a nearly
straight line with the slope of n in different annealing temperatures, the
crystallization volume fractions are chosen from 0.2 to 0.8. Based on
this, the kinetic parameters of the Cu38Zr46Ag8Al8 BMG at different
annealing temperatures in isothermal crystallization conditions can be
calculated, and the values are given in Table 1.
The Avrami exponents vary from 4.10 to 4.74 for different annealing
temperatures and values are listed in Table 1. Considering the average
Avrami exponent n in Cu38Zr46Ag8Al8 BMG is about 4.47 0.33,
suggesting that the crystallization is mainly governed by a constant
nucleation rate [47], similar values have been reported in other BMGs:
Zr55.9Cu18.6Ta8Al7.5Ni10[48], Al89La6Ni5[49] and Ti50Ni25Cu25[50].
3.2.3. Local Avrami exponent
For amorphous materials, the nucleation and growth rates of crystals
do not remain a constant value during the whole crystallization process.
Therefore, the relationship between the local Avrami exponent and the
crystallization volume fraction should be introduced. Based on Eq. (9),
the local Avrami exponent is dened as [51]:
n

ln ln 1 
:
ln t

10

Fig. 13 shows the relationship between the local Avrami exponent


and the crystallization volume fraction at ve different annealing
temperatures, the ve curves display the consistent tendency. According
to the fact that when is between 10% and 20%, n() = 3.04.0, which
means that the crystallization progresses with a decreasing crystallization rate, but the crystallization quantity is increasing in this stage.

Fig. 12. Avrami plots of Cu 38 Zr 46 Ag 8 Al 8 bulk metallic glass at different annealing


temperatures. The Avrami exponents in different annealing temperatures are: 4.74 in
755 K, 4.24 in 761 K, 4.09 in 767 K, 4.54 in 773 K and 4.72 in 779 K, the average value is
4.47. The ts all have the same R2 value of 0.999.

After this stage, the local Avrami exponent is larger than 4.0 which
means that the nucleation rate increases with time. Meanwhile, the
crystallization quantity increased rapidly and the n() value almost
maintains a stable state until the reaches 80%, after that the n()
value is back to 4.0, which means that the nucleation rate decreased to
a constant small value; the main reason is due to the lack of solvend.
After this stage, with the increase of crystallization fraction, the nucleation
rate decreases until the crystallization process is completed. Similar
phenomenon has been reported by Sun et al. [52] in the work of
Fe33Zr67 metallic glass.

4. Conclusion
The crystallization kinetics of Cu38Zr46Ag8Al8 BMG is investigated by
DSC in both non-isothermal and isothermal conditions. The main results
of this work are shown as follows:
In non-isothermal condition, the average crystallization enthalpy is
about 29.36 3.44 J/g, and the average activation energy is about
310 kJ/mol.
In non-isothermal condition, when the crystallization fraction is
between 20% and 80% the Avrami exponent is larger than 4, which
means that the nucleation rate increased with time, and the crystallization quantity increased rapidly. When the crystallization fraction
reaches 80%, the Avrami exponent value is back to 34, which
means that the crystallization progresses with a decreasing crystallization rate, the main reason is due to the lack of solvend.
In isothermal condition, the average crystallization enthalpy is about
1.63 0.12 J/g, and the average value of activation energy is about
451 23 kJ/mol.
In isothermal condition, the Avrami exponent n ranges from 4.10 to
4.74, the crystallization mechanism is mainly governed by constant
nucleation rate.

Table 1
The Avrami exponent n, reaction rate constant K and incubation time at different annealing
temperatures in Cu38Zr46Ag8Al8 bulk metallic glass. The Avrami exponent and reaction rate
constant K are obtained by tting Fig. 12, and the ts all have the same R2 value of 0.999.

Fig. 11. Activation energy of Cu38Zr46Ag8Al8 bulk metallic glass for different crystallization
volume fractions in isothermal heating conditions by tting the experimental points in
Fig. 10. The average activation energy is 451 23 kJ/mol.

Annealing temperature (K)

(min)

755
761
767
773
779

4.74
4.24
4.09
4.54
4.72

3.12
2.48
1.15
0.55
0.12

0.19
0.35
0.50
0.77
1.21

12

J. Cui et al. / Journal of Non-Crystalline Solids 404 (2014) 712

Fig. 13. Relationship between local Avrami exponent n() and crystallization volume
fraction at different annealing temperatures as a guide to the eyes.

Acknowledgments
This work was supported by the Fundamental Research Fund of
Northwestern Polytechnical University (JC20120203) and the Program
of Introducing Talents of Discipline to Universities (B08040). One of
the authors J. Cui wants to acknowledge China Scholarship Council
(CSC) (2011629115) for the nancial support.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]

M.M. Trexler, N.N. Thadhani, Prog. Mater. Sci. 55 (2010) 759.


W.H. Wang, Prog. Mater. Sci. 52 (2007) 540.
A.L. Greer, Science 267 (1995) 1947.
W.H. Wang, C. Dong, C.H. Shek, Mater. Sci. Eng. R 44 (2004) 45.
W.H. Wang, Adv. Mater. 21 (2009) 4524.
C.J. Byrne, M. Eldrup, Science 321 (2008) 502.
Y.J. Huang, J. Shen, J.J. Chen, J.F. Sun, J. Alloys Compd. 447 (2009) 920.
A. Inoue, Acta Mater. 48 (2000) 279.
A. Inoue, W. Zhang, T. Zhang, K. Kurosaka, Mater. Trans. 42 (2001) 1149.
A. Inoue, W. Zhang, Mater. Trans. 43 (2002) 2921.
T. Zhang, A. Inoue, Mater. Trans. 43 (2002) 1367.
J.N. Mei, J.L. Soubeyroux, J.S. Li, J.J. Blandin, H.C. Kou, H.Z. Fu, L. Zhou, J. Non-Cryst.
Solids 357 (2011) 110.
[13] Y.D. Sun, Z.Q. Li, J.S. Liu, J.N. Yang, M.Q. Cong, J. Alloys Compd. 506 (2010) 6.
[14] J. Saida, M. Matsushita, T. Zhang, A. Inoue, M.W. Chen, T. Sakurai, Appl. Phys. Lett. 75
(1999) 3497.

[15] D.H. Xu, G. Duan, W.L. Johnson, Phys. Rev. Lett. 92 (2004) 245504.
[16] S. Venkataraman, H. Hermann, D.J. Sordelet, J. Eckert, J. Appl. Phys. 104 (2008)
066107.
[17] S. Venkataraman, E. Rozhkova, J. Eckert, L. Schulta, D.J. Sordelet, Intermetallics 13
(2005) 833.
[18] J.C. Qiao, J.M. Pelletier, J. Mater. Sci. Technol. 30 (2014) 523.
[19] H.E. Kissinger, Anal. Chem. 29 (1957) 1702.
[20] H.L. Friedman, J. Ploym. Sci. Symp. 6 (1964) 183.
[21] J. estk, J. Therm. Anal. Calorim. 110 (2012) 5.
[22] J. estk, J. Therm. Anal. 36 (1991) 1997.
[23] R.T. Savalia, K.N. Lad, A. Pratap, G.K. Dey, S. Banerjee, J. Therm. Anal. Calorim. 78
(2004) 745.
[24] F. Xu, Y.L. Du, P. Gao, Z.D. Han, G. Chen, S.Q. Wang, J.Z. Jiang, J. Alloys Compd. 441
(2007) 76.
[25] X.F. Wu, L.K. Meng, W. Zhao, Z.Y. Suo, Y. Si, K.Q. Qiu, J. Rare Earths 25 (2007) 189.
[26] D. Okai, Y. Shimizu, N. Hirano, T. Fukami, T. Yamasaki, A. Inoue, J. Alloys Compd. 504
(2010) S247.
[27] J.T. Zhang, W.M. Wang, H.J. Ma, G.H. Li, R. Li, Z.H. Zhang, Thermochim. Acta 505
(2010) 41.
[28] G.Q. Xie, D.V. Louzguine-Luzgin, Q.S. Zhang, W. Zhang, A. Inoue, J. Alloys Compd. 483
(2009) 24.
[29] J.M. Pelletier, B. Van de Moortle, J. Non-Cryst. Solids 325 (2003) 133.
[30] Y.J. Yang, D.W. Xing, J. Shen, J.F. Sun, S.D. Wei, H.J. He, D.G. McCartney, J. Alloys
Compd. 415 (2006) 106.
[31] S.W. Wei, B.Z. Ding, T.Q. Lei, Mater. Lett. 37 (1998) 265.
[32] X.F. Lu, J.N. Hay, Polymer 42 (2001) 9423.
[33] Y. Liu, G.S. Yang, Thermochim. Acta 500 (2010) 13.
[34] S.H. Kim, A.H. Ahh, T. Hirai, Polymer 44 (2003) 5625.
[35] J. estk, G. Berggren, Thermochim. Acta 3 (1971) 1.
[36] K.N. Lad, R.T. Savalia, A. Pratap, G.K. Dey, S. Banerjee, Thermochim. Acta 473 (2008)
74.
[37] N. Mehta, K. Singh, N.S. Saxena, Phys. B 403 (2008) 3928.
[38] A.A. Soliman, S. Al-Heniti, A. Al-Hajry, M. Al-Assiri, G. Al-Barakati, Thermochim. Acta
413 (2004) 57.
[39] W. Lu, B. Yan, W.H. Huang, J. Non-Cryst. Solids 351 (2005) 3320.
[40] J.S.C. Jang, L.J. Chang, G.L. Chen, J.C. Huang, Intermetallics 13 (2005) 907.
[41] J. Wang, H.C. Kou, J.S. Li, X.F. Gu, H. Zhong, H. Chang, L. Zhou, J. Phys. Chem. Solids 70
(2009) 1448.
[42] D. Kasase, A.P. Tsai, A. Inoue, T. Masumoto, Appl. Phys. Lett. 62 (1993) 137.
[43] J.C. Qiao, J.M. Pelletier, J. Non-Cryst. Solids 357 (2011) 2590.
[44] J.Z. Jiang, Y.X. Zhuang, H. Rasmussen, J. Saida, A. Inoue, Phys. Rev. B 64 (2001)
094208.
[45] Z.Z. Yuan, X.D. Chen, B.X. Wang, Y.J. Wang, J. Alloys Compd. 407 (2006) 163.
[46] F. Xu, J.Z. Jiang, Q.P. Cao, Y.W. Du, J. Alloys Compd. 392 (2005) 173.
[47] F. Liu, F. Sommer, C. Bos, E.J. Mittemeijer, Int. Mater. Rev. 52 (2007) 193.
[48] Q. Chen, L. Liu, K.C. Chan, J. Alloys. Compd. 419 (2006) 71.
[49] F. Ye, K. Lu, J. Non-Cryst. Solids 262 (2000) 228.
[50] D.V. Louzguine, A. Inoue, J. Mater. Sci. 35 (2000) 4159.
[51] Y.L. Gao, J. Shen, J.F. Sun, G. Wang, D.W. Xing, H.Z. Xian, B.D. Zhou, Mater. Lett. 57
(2003) 1894.
[52] N.X. Sun, X.D. Liu, K. Lu, Scr. Mater. 34 (1996) 1201.

Anda mungkin juga menyukai