kinetic analysis of the non-isothermal crystallization ... · 0.5 bulk metallic glasses were...

10
Kinetic analysis of the non-isothermal crystallization process, magnetic and mechanical properties of FeCoBSiNb and FeCoBSiNbCu bulk metallic glasses Parthiban Ramasamy, Mihai Stoica, A. H. Taghvaei, K. G. Prashanth, Ravi Kumar, and Jürgen Eckert Citation: Journal of Applied Physics 119, 073908 (2016); doi: 10.1063/1.4942179 View online: http://dx.doi.org/10.1063/1.4942179 View Table of Contents: http://scitation.aip.org/content/aip/journal/jap/119/7?ver=pdfcov Published by the AIP Publishing Articles you may be interested in Enhanced glass-forming ability of FeCoBSiNb bulk glassy alloys prepared using commercial raw materials through the optimization of Nb content J. Appl. Phys. 107, 09A315 (2010); 10.1063/1.3350898 Two-stage-like glass transition and the glass-forming ability of a soft magnetic Fe-based glassy alloy J. Appl. Phys. 105, 053518 (2009); 10.1063/1.3080139 Fe-based bulk glassy alloy composite containing in situ formed α - ( Fe , Co ) and ( Fe , Co ) 23 B 6 microcrystalline grains Appl. Phys. Lett. 89, 101915 (2006); 10.1063/1.2348737 Influence of Ni on the structural and magnetic properties of Ni x Fe 73.5 − x Si 13.5 B 9 Nb 3 Cu 1 ( 0 x 25 ) alloys J. Appl. Phys. 97, 023901 (2005); 10.1063/1.1825633 Superhigh strength and good soft-magnetic properties of ( Fe , Co ) – B – Si – Nb bulk glassy alloys with high glass-forming ability Appl. Phys. Lett. 85, 4911 (2004); 10.1063/1.1827349 Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. IP: 147.171.56.188 On: Fri, 19 Feb 2016 16:46:34

Upload: others

Post on 04-Jul-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Kinetic analysis of the non-isothermal crystallization ... · 0.5 bulk metallic glasses were evaluated using differential scanning calorimetry under non-isothermal condition. The

Kinetic analysis of the non-isothermal crystallization process, magnetic andmechanical properties of FeCoBSiNb and FeCoBSiNbCu bulk metallic glassesParthiban Ramasamy, Mihai Stoica, A. H. Taghvaei, K. G. Prashanth, Ravi Kumar, and Jürgen Eckert Citation: Journal of Applied Physics 119, 073908 (2016); doi: 10.1063/1.4942179 View online: http://dx.doi.org/10.1063/1.4942179 View Table of Contents: http://scitation.aip.org/content/aip/journal/jap/119/7?ver=pdfcov Published by the AIP Publishing Articles you may be interested in Enhanced glass-forming ability of FeCoBSiNb bulk glassy alloys prepared using commercial raw materialsthrough the optimization of Nb content J. Appl. Phys. 107, 09A315 (2010); 10.1063/1.3350898 Two-stage-like glass transition and the glass-forming ability of a soft magnetic Fe-based glassy alloy J. Appl. Phys. 105, 053518 (2009); 10.1063/1.3080139 Fe-based bulk glassy alloy composite containing in situ formed α - ( Fe , Co ) and ( Fe , Co ) 23 B 6microcrystalline grains Appl. Phys. Lett. 89, 101915 (2006); 10.1063/1.2348737 Influence of Ni on the structural and magnetic properties of Ni x Fe 73.5 − x Si 13.5 B 9 Nb 3 Cu 1 ( 0 x 25 )alloys J. Appl. Phys. 97, 023901 (2005); 10.1063/1.1825633 Superhigh strength and good soft-magnetic properties of ( Fe , Co ) – B – Si – Nb bulk glassy alloys with highglass-forming ability Appl. Phys. Lett. 85, 4911 (2004); 10.1063/1.1827349

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. IP: 147.171.56.188 On: Fri, 19 Feb 2016 16:46:34

Page 2: Kinetic analysis of the non-isothermal crystallization ... · 0.5 bulk metallic glasses were evaluated using differential scanning calorimetry under non-isothermal condition. The

Kinetic analysis of the non-isothermal crystallization process, magnetic andmechanical properties of FeCoBSiNb and FeCoBSiNbCu bulk metallicglasses

Parthiban Ramasamy,1,a) Mihai Stoica,1,2 A. H. Taghvaei,3 K. G. Prashanth,4 Ravi Kumar,5

and J€urgen Eckert6,7

1IFW Dresden, Institute for Complex Materials, Helmholtzstr. 20, D-01069 Dresden, Germany2Politehnica University Timisoara, P-ta Victoriei 2, RO-300006 Timisoara, Romania3Department of Materials Science and Engineering, Shiraz University of Technology, Shiraz, Iran4Additive Manufacturing Center, Sandvik AB, 81181 Sandviken, Sweden5Department of Metallurgical and Materials Engineering, Indian Institute of Technology Madras,600036 Chennai, India6Erich Schmid Institute of Materials Science, Austrian Academy of Sciences, Jahnstraße 12,A-8700 Leoben, Austria7Department Materials Physics, Montanuniversit€at Leoben, Jahnstraße 12, A-8700 Leoben, Austria

(Received 7 October 2015; accepted 1 February 2016; published online 19 February 2016)

The crystallization kinetics of [(Fe0.5Co0.5)0.75B0.2Si0.05]96Nb4 and {[(Fe0.5Co0.5)0.75B0.2Si0.05]0.96

Nb0.04}99.5Cu0.5 bulk metallic glasses were evaluated using differential scanning calorimetry under

non-isothermal condition. The fully glassy rods with diameters up to 2 mm were obtained by copper

mold injection casting. Both glasses show good thermal stability, but the addition of only 0.5% Cu

completely changes the crystallization behavior. The average activation energy required for crystalliza-

tion decreases from 645 kJ/mol to 425 kJ/mol after Cu addition. Upon heating, the Cu-free alloy forms

only the metastable Fe23B6 phase. In contrast, two well-separated exothermic events are observed for

the Cu-added bulk glassy samples. First, the (Fe,Co) phase nucleates and then (Fe,Co)2B and/or

(Fe,Co)3B crystallize from the remaining glassy matrix. The Cu-added alloy exhibits a lower coercivity

and a higher magnetic saturation than the base alloy, both in as-cast as well as in annealed condition.

Besides, the Cu-added glassy sample with 2 mm diameter exhibits a maximum compressive fracture

strength of 3913 MPa together with a plastic strain of 0.6%, which is highest plastic strain ever reported

for 2 mm diameter ferromagnetic bulk metallic glass sample. Although Cu addition improves the mag-

netic and mechanical properties of the glass, it affects the glass-forming ability of the base alloy.VC 2016 AIP Publishing LLC. [http://dx.doi.org/10.1063/1.4942179]

I. INTRODUCTION

The first successful synthesis of Fe-based bulk metallic

glasses (BMGs) by copper mold casting was reported by

Inoue et al. in 1995, for the Fe-Al-Ga-P-C-B system.1 Since

then, various kinds of Fe-based BMGs have been produced

using the copper mold casting method. Some of these alloys

are designed for application as functional2,3 and structural

materials,4,5 not only due to their good soft-magnetic proper-

ties but also because of their respective high fracture strength

combined with good glass-forming ability (GFA). These

Fe-based BMGs do not exhibit strain hardening and can sus-

tain almost no plastic deformation. This brittle behavior

under mechanical loading severely limits the applications of

Fe-based BMGs as engineering materials. Great efforts have

been made recently for improving the plastic deformability

of monolithic BMGs or BMG composites. For example, the

compressive deformability has been largely enhanced in Zr-,6

Cu-,7 Ni-,8 and Pd-based9 BMG composites through the

introduction of second phase particles, fine pores, or notches

in the glassy matrix.10,11 In the case of Fe-based BMGs, the

plasticity can be improved by adding small amounts of soft

metals like Ni or Er.9,12

In the last decade, (Fe-Co)-B-Si-Nb BMGs have been

regarded as one of the most promising Fe-based glassy alloys

combining the advantages of functional and structural materi-

als.13 Yoshizawa et al. designed the FINEMET alloys by add-

ing a small amount of copper into an Fe-based metallic glass

and obtained excellent soft magnetic properties upon anneal-

ing.14 Shen et al. revealed that in situ formation of (Fe, Co)

and (Fe, Co)23B6 microcrystalline grains during the solidifica-

tion process can improve the ductility of {[(Fe0.5Co0.5)0.75

Si0.05B0.20]0.96Nb0.0499.75Cu0.25 BMG composites.15 The

effect of crystallization upon heat treatment of Fe71.5�xCox

B13.5Si10Nb4Cu1 glassy alloys on their soft magnetic proper-

ties was reported by Inoue and Shen.16 For the [(Fe0.5

Co0.5)0.75B0.2Si0.05]96Nb4 BMG, the thermal stability, in-situcrystallization, and crystallization kinetics were reported by

Stoica et al.17 In general, the Johnson–Mehl–Avrami–

Kolmogorov (JMAK) approach is used to analyze the crystal-

lization kinetics under isothermal conditions.18–20 However,

several researchers also pointed out that the JMAK equation

can be applied to non-isothermal experiments, when the entire

nucleation process occurs during the early crystallization

stages and becomes negligible in the later stage as well as

a)Author to whom correspondence should be addressed. Electronic mail:

[email protected]

0021-8979/2016/119(7)/073908/9/$30.00 VC 2016 AIP Publishing LLC119, 073908-1

JOURNAL OF APPLIED PHYSICS 119, 073908 (2016)

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. IP: 147.171.56.188 On: Fri, 19 Feb 2016 16:46:34

Page 3: Kinetic analysis of the non-isothermal crystallization ... · 0.5 bulk metallic glasses were evaluated using differential scanning calorimetry under non-isothermal condition. The

precipitated phases uniformly distribute in the amorphous

matrix.21 In such a case, the precipitated phases nucleation

will not proceed in parallel with crystal growth, the crystalli-

zation rate is controlled by the temperature and independent

of the previous thermal process.21,22 Based on this considera-

tion, the crystallization kinetics and the corresponding crystal-

lization mechanism of several systems have been discussed

successfully for non-isothermal conditions using the JMAK

equation.22–25 Bl�azquez et al.26 have pointed that the non-

isothermal experiments are relatively easier and have a lower

noise-to-signal ratio for kinetic experiments.

In this paper, [(Fe0.5Co0.5)0.75B0.2Si0.05]96Nb4 (named

further “base alloy” or BA) and {[(Fe0.5Co0.5)0.75 Si0.05

B0.20]0.96Nb0.04}99.5Cu0.5 (named further “modified alloy” or

MA) glassy alloys were chosen for investigation. They have

very attractive structural and functional properties, as, for

example, a fracture strength rf of �4000 MPa, a saturation

magnetization Ms of �1T, a relative permeability l of

�12,000, as well as a coercivity Hc of 2 A/m.13,27–29 The

thermal stability, phase evaluation, kinetic parameters, and

magnetic properties of these glassy alloys were evaluated.

The results give more details to understand the relationship

between the transformation of the metastable glassy phase

and the improvement of the soft magnetic properties upon

Cu addition and they are discussed in comparison with the

parent alloy [(Fe0.5Co0.5)0.75B0.2Si0.05]96Nb4.

II. EXPERIMENTAL DETAILS

The BA and MA master alloys were prepared in several

steps. In a first step, a eutectic Fe25Nb75 (wt. %) prealloy was

produced by arc-melting pure Fe (99.9 wt. %) and Nb

(99.9 wt. %) lumps. Then, the FeNb prealloy, together with

the rest of necessary Fe and Co lumps (99.9 wt. %), crystal-

line B (99 wt. %), and crystalline Si (99.99 wt. %) were

melted together by induction heating under a protective Ar

atmosphere. In this process, induction melting was preferred

in order to assure a good homogeneity of the entire master

alloy. The modified MA alloy was produced by adding pure

Cu (99.99%) to the BA by arc-melting. Pieces of the master

alloys were remelted in quartz tubes and subsequently

injected into a water-cooled copper mold under a high purity

Ar atmosphere to produce rod-shaped specimens with a

maximum diameter of 2 mm and a length of 60 mm. The

thermal behavior of the glassy samples was evaluated with a

NETZSCH DSC 404 C differential scanning calorimeter

(DSC). For this purpose, the specimens were prepared from

the as-cast rods and heated in Al2O3 crucibles. The activation

energies for the obtained glassy rods were evaluated using

the Kissinger method,30 with a set of heating rates ranging

from 5 to 20 K/min. Structural characterization of the as-cast

rods and the crystallized samples was performed by X-ray

diffraction (XRD) using a PHILIPS PW 1050 diffractometer

with Co-Ka radiation (k¼ 0.17889 nm). The samples were

annealed for 5 min at different temperatures using the same

NETZSCH DSC 404C, with a constant heating and cooling

rate of 20 K/min. For magnetic measurements, DC M-H hys-

teresis loops were measured with a vibrating sample magne-

tometer (VSM) at ambient temperature. The coercivity was

measured using a Foerster Coercimat-type device under a

DC magnetic field, which can be continuously changed from

�250 mT to þ250 mT. Cylindrical specimens of 2 mm diam-

eter and 4 mm length were prepared from the as-cast samples

and tested with an INSTRON 8562 testing facility in con-

stant strain rate mode (strain rate¼ 1� 10�4 s�1) at room

temperature. Both ends of the samples were carefully pol-

ished to make them parallel to each other prior to the com-

pression test. The density of the samples was evaluated by

the Archimedes method using a computer controlled micro-

balance and dodecane (C12H26) as working liquid. The final

values were obtained by averaging over 25 experimental val-

ues. Hardness measurements were done using a computer

controlled STRUERS DURAMIN 5 Vickers hardness tester

with an applied load of 0.98 N for 10 s. The final hardness

values were obtained after averaging 10 experimental data

points. The accuracy of the measured data lies within

62.5 K in the case of the DSC measurements, 60.1 A/m for

the coercivity, 680 A/m (�1 Oe) for the VSM measure-

ments, 0.5% for the density measurements, 610 HV for the

hardness measurements, and 620 MPa for the stress-strain

measurements.

III. RESULTS AND DISCUSSION

Following the experimental route as described earlier,

the maximum size of the fully glassy rods obtained for the

{[(Fe0.5Co0.5)0.75B0.2Si0.05]0.96Nb0.04}99.5Cu0.5 alloy is 2 mm

in diameter and 60 mm in length. For a similar alloy, with

0.6 at. % Cu, Jia et al.31 reported the formation of fully

glassy rods with various cross-sections (i.e., 2, 3, and 4 mm

diameter), each part extending only over 1–2 cm. Therefore,

the actual samples (i.e., 2 mm in diameter and 60 mm in

length) seem to be the largest fully glassy {[(Fe0.5Co0.5)0.75

B0.2Si0.05]0.96Nb0.04}99.5Cu0.5 cast rods.

A. Thermal analysis and structural evolution

Figures 1(a) and 1(b) show the DSC traces measured at

different heating rates of 5, 10, 15, and 20 K/min for the

as-cast BA and MA samples, respectively. The temperatures

Tg, Tx, and Tp, which are marked in the DSC curves, are the

glass transition temperature, the onset temperature of crystal-

lization, and the crystallization peak temperature, respec-

tively. The BA shows one main exothermic event, whereas

the MA exhibits two exothermic events. The addition of Cu

completely changes the crystallization behavior (e.g., at a

heating rate of 20 K/min, the BA exhibits a single exother-

mic event at 867 K, whereas the MA shows two exothermic

peaks at 838 (Tp1) and 945 K (Tp2), respectively). In case of

the MA, the primary crystallization process shifted towards

lower temperatures and extends over a range of 100 K,

whereas the secondary crystallization gets completed within

a smaller temperature range. Regardless of the heating rate,

all curves clearly exhibit a distinct Tg and a supercooled liq-

uid region (SLR) followed by crystallization. From the DSC

curves, it is evident that all the characteristic temperatures

(Tg, Tx, Tp1, and Tp2) shift to higher values with increasing

heating rate (as shown in Table I). The extension of the SLR

073908-2 Ramasamy et al. J. Appl. Phys. 119, 073908 (2016)

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. IP: 147.171.56.188 On: Fri, 19 Feb 2016 16:46:34

Page 4: Kinetic analysis of the non-isothermal crystallization ... · 0.5 bulk metallic glasses were evaluated using differential scanning calorimetry under non-isothermal condition. The

(i.e., DTx¼ Tx -Tg) is larger at higher heating rates indicating

that Tx is more sensitive to the heating rate than Tg.

The structural evolution of the MA at different tempera-

tures (i.e., 818, 903, and 978 K) was studied using XRD

(shown in Fig. 2). After heating to the first crystallization

temperature (Tx¼ 818 K), bcc-(Fe,Co) crystals starts to nu-

cleate. The existence of a residual glassy matrix is confirmed

by a broad diffraction maximum in the XRD pattern of the

sample annealed at 818 K. The nucleation of the bcc-(Fe,Co)

crystals at lower temperatures may be attributed to the pres-

ence of a small amount of Cu atoms in the matrix (i.e., Fe

and Cu have a positive heat of mixing), which can accelerate

the initialization of the nanocrystals.14 However, according

to Fig. 1(b), Cu addition extends the time and temperature

required for completion of crystallization. At 902 K, where

the growth of (Fe,CO) is almost close to be complete, a sec-

ond glass transition event is observed (i.e., an event associ-

ated most probably with the glass transition of the remaining

glassy matrix). Further, at 973 K, i.e., at the end of the sec-

ond crystallization step, the pattern reveals the formation of

the tetragonal (Fe,Co)2B phase and metastable orthorhombic

(Fe,Co)23B6. These results are in agreement with those

recently presented in Ref. 32, which indicates a completely

different crystallization behavior of MA compared to BA. In

the case of BA, as can be seen also in Fig. 1(a), there is only

one crystallization peak in the depicted temperature range,

corresponding to a primary crystallization, where only the

complex (Fe, Co)23B6 phase forms.15,17,33

The apparent activation energy required for each charac-

teristic event (Tg, Tx, and Tp) can be calculated using the

Kissinger method:30

lnbT2¼ � E

RT

� �þ constant; (1)

where b is the heating rate, T is the characteristic tempera-

ture at that specific heating rate, E is the activation energy,

and R is the universal gas constant. By plotting -ln(b/T2) as a

function of 1000/T, a linear dependence with slope E/R is

obtained (as presented in Fig. 3). From this slope, the activa-

tion energy for each characteristic event was calculated and

the data are given in Table II. The activation energy for the

glass transition in BA and MA are Eg¼ 582 6 15 and

536 6 14 kJ/mol, respectively. The lower activation energy

in case of MA is due to the addition of Cu atoms, which

reduces the energy barrier required to reconstruct the atomic

configuration.34 Recently, Stoica et al.32,35 studied the struc-

ture of as-cast rods as well as their in-situ relaxation and

crystallization behavior of both BA and MA specimens. In

their studies, they revealed that the presence of copper atoms

FIG. 1. Isochronal DSC curves recorded at different heating rates for

(a) [(Fe0.5Co0.5)0.75B0.2Si0.05]96Nb4 and (b) {[(Fe0.5Co0.5)0.75B0.2Si0.05]0.96

Nb0.04}99.5Cu0.5 glassy samples.

TABLE I. Glass transition (Tg), onset (Tx), crystallization peak (Tp1 and Tp2)

temperatures, and the extension of the supercooled liquid region (DTx) for

the [(Fe0.5Co0.5)0.75B0.2Si0.05]96Nb4 (BA) and {[(Fe0.5Co0.5)0.75B0.2Si0.05]0.96

Nb0.04}99.5Cu0.5 (MA) glasses. The accuracy of the experimental data lies

within 62.5 K.

5 (K/min) 10 (K/min) 15 (K/min) 20 (K/min)

BA MA BA MA BA MA BA MA

Tg (K) 812 780 817 786 821 790 825 793

Tx (K) 837 797 843 804 849 811 858 818

Tp1(K) 853 818 860 828 863 834 867 838

Tp2 (K) … 933 … 939 … 941 … 945

DTx (K) 25 17 26 18 28 21 33 25

FIG. 2. XRD patterns of the {[(Fe0.5Co0.5)0.75B0.2Si0.05]0.96Nb0.04}99.5Cu0.5

glass in the as-cast state and after annealing at different temperatures.

073908-3 Ramasamy et al. J. Appl. Phys. 119, 073908 (2016)

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. IP: 147.171.56.188 On: Fri, 19 Feb 2016 16:46:34

Page 5: Kinetic analysis of the non-isothermal crystallization ... · 0.5 bulk metallic glasses were evaluated using differential scanning calorimetry under non-isothermal condition. The

in the MA may lead to the formation of medium range order

(MRO) clusters, which in turn leads to an increased compres-

sive plastic deformability and also promotes two glass

transition-like events. The activation energy required for the

nucleation Ex and the first crystallization Ep1 for the MA is less

than the Ex and Ep1 values for the BA (see Table II). This could

be due to the presence of MRO clusters present in the as-cast

alloy. These clusters act as a heterogeneous nucleation sites,

which lowers the energy required for nucleation. The higher

Ep2 value for the MA indicates that the second crystallization

event is more difficult to occur compared to the first one.

B. Crystallization kinetics

From the isochronal DSC curves, the crystallized volume

fraction for the non-isothermal crystallization can be deduced

as function of temperature using the following equation:36,37

x ¼

ðT

T0

dHc=dTð ÞdT

ðT1

T0sdHc=dTð ÞdT

¼ A

A0

; (2)

where T0 and T1 are the temperatures corresponding to the

onset and end of crystallization and dHc/dT is the heat

capacity at constant pressure. A is the area under the DSC

curve between the onset temperature and a given tempera-

ture, and A0 is the area between onset and end of crystalliza-

tion. The crystallized volume fraction is plotted as a function

of temperature for both BA and MA in Figs. 4(a) and 4(b),

respectively. Apparently, the variation of the crystallized

volume fraction versus temperature follows a sigmoidal

behavior. The sigmoidal plots indicate that the crystallization

occurs through bulk growth of nuclei rather than the growth

of pre-existing crystals.38 Minor variations in the slope of

the crystallized volume fraction curves for both alloys BA

and MA at the beginning (i.e., x< 0.1) and the end (i.e.,

x> 0.9) of the process indicate a slow reaction rate at these

stages. In general, the crystallization with sigmoidal curves

can be divided into three stages. In first stage, the nuclei pre-

cipitate at various points in the sample and the nucleation

becomes dominant. At the second stage, the increase in the

surface area between the nuclei and the amorphous matrix

leads to a drastic increase in the crystallized volume fraction

and, consequently, stable crystallization occurs. In the last

stage, the crystallization process slows down as a result of

crystal impingement and the decrease in the surface area

between the crystallized phase and the residual amorphous

matrix. Figure 5 shows the relation between the crystallized

volume fraction (x) and the local activation energy Ea for

both alloys in the whole crystallization process. The activa-

tion energy required for the completion of the crystallization

FIG. 3. Kissinger plot for the calculation of activation energies required by

different events in the case of {[(Fe0.5Co0.5)0.75B0.2Si0.05]0.96Nb0.04}99.5Cu0.5

glass.

TABLE II. Apparent activation energies (Eg, Ex, Ep1, Ep2) for the

[(Fe0.5Co0.5)0.75B0.2Si0.05]96Nb4, {[(Fe0.5Co0.5)0.75B0.2Si0.05]0.96Nb0.04}99.5Cu0.5

glasses.

Alloy Eg (kJ/mol) Ex (kJ/mol) Ep1 (kJ/mol) Ep2 (kJ/mol)

FeCoBSiNb 582 6 15 370 6 6 607 6 15 …

FeCoBSiNbCu 536 6 14 340 6 6 374 6 8 748 6 19

FIG. 4. Crystallized volume fractions(x) at different heating rate for (a)

[(Fe0.5Co0.5)0.75B0.2Si0.05]96Nb4 and (b) {[(Fe0.5Co0.5)0.75B0.2Si0.05]0.96Nb0.04}99.5

Cu0.5 glassy samples.

073908-4 Ramasamy et al. J. Appl. Phys. 119, 073908 (2016)

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. IP: 147.171.56.188 On: Fri, 19 Feb 2016 16:46:34

Page 6: Kinetic analysis of the non-isothermal crystallization ... · 0.5 bulk metallic glasses were evaluated using differential scanning calorimetry under non-isothermal condition. The

event increases gradually with increasing crystallized vol-

ume fraction, indicating that a larger energy barrier is

required for the growth of the (Fe,Co) crystals compared to

their nucleation. During the growth of the (Fe,Co) phase,

the composition of the residual matrix changes continu-

ously and it becomes enriched in Nb, as well as in B

and Si. Such change in composition of the remaining amor-

phous phase may increase its stability against crystalliza-

tion, thus leading to a higher activation energy. In case

of the BA, the activation energy increases relatively

steeper than for the MA. This effect can be due to the pres-

ence of fewer nucleation sites in the BA than in the MA.

The average activation energy required for the completion

of the whole crystallization process in case of the MA is

425 kJ/mol, which is lower than the BA average activation

energy of 645 kJ/mol (see Fig. 5).

1. Extension to non-isothermal processes

The nucleation and growth rates of crystals do not

remain constant throughout the crystallization process of an

amorphous alloy. To describe the crystallization mechanism

at every step, the local Avrami exponent, n(x), was intro-

duced.18–20 The isothermal crystallization kinetics of metal-

lic glasses is usually studied using the JMAK theory,18–20,39

which can be expressed by the JMAK equation as follows:

x ¼ 1� exp½�fkðt� t0Þgn�; (3)

where k is the frequency factor, t is the time, to is the incuba-

tion time, n is the Avrami exponent, and x is the effective

crystalline volume fraction. The JMKA equation is often

written as

ln½�lnð1� xÞ� ¼ n lnðt� t0Þ þ const: (4)

For isothermal processes, the local Avrami exponent n(x)can be calculated by differentiating the JMAK equation (i.e.,

Eq. (4))

n xð Þ ¼ d ln �ln 1� xð Þ½ �d ln t� t0ð Þ

: (5)

For non-isothermal processes, Nakamura et al.36,37 gener-

alized the JMAK equation under the approximation of

isokinetic behavior. This approximation implies that the

crystallization process is independent of the thermal history

of the sample. The modified JMAK equation used to calcu-

late the local Avrami exponent for a non-isothermal pro-

cess can be written as follows:26

n xð Þ ¼ 1

1þ E

RT1� T0

T

� � dln �ln 1� xð Þ½ �

d lnT � T0

b

� �� � ; (6)

where T0 is the onset temperature of crystallization, b is the

heating rate, and E is the activation energy. The plots of

ln½�lnð1� xÞ� versus fln½T�T0

b �g for the BA and MA samples

are shown in Figs. 6(a) and 6(b), respectively. The slopes of

the curves vary in different crystallization stages, which indi-

cate that the nucleation and growth do not remain the same

during the whole crystallization. The slope of the KJMA plots

for BA changes slightly in the initial stage, while the changes

become faster in the later stage. In addition, it can be seen that

the rate of the slope changes increases with increasing heating

rate. In case of the MA, the slope decreases uniformly at lower

heating rates, and it gradually increases. The local Avrami

exponent n(x) can be used to characterize the nucleation and

growth behavior with increasing crystallized volume fraction.

The local Avrami exponent plots (i.e., the results calculated

using Eq. (6)) for the BA and MA specimens are shown in the

insets of Figs. 6(a) and 6(b), respectively. The Avrami expo-

nent which reflects the nucleation and growth parameters can

be expressed as:40,41

n ¼ aþ bp; (7)

where a is the nucleation index (a¼ 0 for zero nucleation

rate, a< 0< 1 for decreasing nucleation rate, a¼ 1 for con-

stant nucleation rate, and a> 1 for increasing nucleation

rate), b is the dimensionality of the growth (i.e., 1, 2, or 3

dimensional growth), and p is the growth index (p¼ 0.5 for

diffusion-controlled and p¼ 1 for interface-controlled

growth). To compare the crystallization kinetics between BA

and MA, the variation of the local Avrami exponent as a

function of crystallized volume fraction (n(x) vs x) at a heat-

ing rate of 20 K/min is presented in Fig. 7. It has been

reported that the crystallization in Fe-based metallic glasses

mostly occurs by a diffusion controlled process (p¼ 0.5 in

Eq. (7)).17,42,43 According to the diffusion-controlled phase

transformation theory,44,45 n< 1.5 demonstrates the growth

of pre-existing nuclei, n¼ 1.5 indicates site saturation or the

growth of the nuclei occurs with zero nucleation rate,

1.5< n< 2.5 implies the growth of precipitates with decreas-

ing nucleation rate, n¼ 2.5 indicates the growth with con-

stant nucleation rate, and n> 2.5 is characteristic of the

growth with increase in nucleation rate. For BA (see Fig. 8),

the local Avrami exponents at all heating rates rapidly

increase to about 1.75 at the beginning of the transformation.

FIG. 5. Local activation energy as a function of crystallized volume fraction

for [(Fe0.5Co0.5)0.75B0.2Si0.05]96Nb4 and {[(Fe0.5Co0.5)0.75B0.2Si0.05]0.96Nb0.04}9.5

Cu0.5 glasses.

073908-5 Ramasamy et al. J. Appl. Phys. 119, 073908 (2016)

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. IP: 147.171.56.188 On: Fri, 19 Feb 2016 16:46:34

Page 7: Kinetic analysis of the non-isothermal crystallization ... · 0.5 bulk metallic glasses were evaluated using differential scanning calorimetry under non-isothermal condition. The

In the next stage, the Avrami exponents at relatively low

heating rates (5, 10, and 15 K/min) start to decrease gradu-

ally as a result of a decrease in the nucleation rate.

Conversely, the Avrami exponent at 20 K/min increases until

reaching a crystallized volume fraction of x¼ 0.1 and subse-

quently decreases. In contrast, for the MA the local Avrami

exponent at higher heating rates (15 and 20 K/min) and lower

heating rates (5 and 10 K/min) increase rapidly to about 2.5

and 2, respectively; and then significantly decrease until the

crystallized volume fraction reaches x¼ 0.9, and finally the

local Avrami exponents starts to increase. Concerning the

BA (see Fig. 7), the Avrami exponent reaches a maximum

value of 2.1 at x¼ 0.1, indicating three-dimensional growth

with decreasing nucleation rate. The gradual decrease of the

Avrami exponent occurs beyond x¼ 0.1, indicating a

decrease in the nucleation rate. As can be seen, the Avrami

exponents are smaller than 1.5 when the crystalized fraction

becomes larger than 0.35, which manifests that further trans-

formation results in growth of the crystals formed upon the

first stage of devitrification without any nucleation. For the

same alloy it was reported17 that the Avrami exponent is

1.43 under isothermal transformation conditions, corre-

sponding to a diffusion-controlled mechanism. In contrast,

for the MA the local Avrami exponents decrease continu-

ously to values below n¼ 1.5 for x> 0.1, indicating that a

faster site saturation occurred compared to the BA. In other

words, the crystallization of the MA is dominantly governed

by three-dimensional growth of pre-existing nuclei (or clus-

ters) formed before crystallization due to the Cu addition.

C. Magnetic properties

The influence of Cu on the behavior of BA was further

examined through investigation of the magnetic properties of

the as-cast and annealed MA samples. In order to avoid the

FIG. 6. Plots of ln½�lnð1� xÞ� vs fln½T � T0=b�g at different heating

rates for (a) [(Fe0.5Co0.5)0.75B0.2Si0.05]96Nb4 and (b) {[(Fe0.5Co0.5)0.75B0.2

Si0.05]0.96Nb0.04}99.5Cu0.5 glassy samples. The inset shows the corresponding

local Avrami exponent n(x) as a function of crystallized volume fraction x,

calculated for different heating rates.

FIG. 7. Combined local Avrami exponent n(x) vs crystallized volume fraction

x of the [(Fe0.5Co0.5)0.75B0.2Si0.05]96Nb4 and {[(Fe0.5Co0.5)0.75B0.2Si0.05]0.96

Nb0.04}99.5Cu0.5 glasses at a heating rate of 20 K/min.

FIG. 8. Hysteresis loops for the {[(Fe0.5Co0.5)0.75B0.2Si0.05]0.96Nb0.04}99.5

Cu0.5 glass in as-cast and annealed condition.

073908-6 Ramasamy et al. J. Appl. Phys. 119, 073908 (2016)

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. IP: 147.171.56.188 On: Fri, 19 Feb 2016 16:46:34

Page 8: Kinetic analysis of the non-isothermal crystallization ... · 0.5 bulk metallic glasses were evaluated using differential scanning calorimetry under non-isothermal condition. The

systematic errors introduced by different sample dimensions

and annealing times, all the sample sizes and annealing times

were kept the same for all the experiments. Table III gives

the coercivity values for the as-cast and annealed MA

samples (heated at a constant rate of 20 K/min in the DSC) at

different temperatures. The as-cast MA samples have a coer-

civity of 2 A/m, which is lower than that of the as-cast BA

(2.7 A/m (Ref. 17)). The coercivity of the MA decreases ini-

tially from 2 A/m for the as-cast state to 1.5 A/m for the sam-

ples annealed at 813 K (i.e., Tx – 5Kof MA) and further

increases when the annealing temperatures are above Tx. The

coercivity finally reaches 37.5 A/m after annealing up to the

end of the first DSC peak (T¼ 918 K). The initial decrease in

coercivity may be due to stress relaxation and a decreasing

density of quasi-dislocation dipole-type (QDD) defects pin-

ning the magnetic domain walls in the amorphous phase.46

On the other hand, the increase of coercivity at higher tem-

peratures can be attributed to the formation of (Fe,Co) crys-

tals in the glassy matrix. It is worth noting that even after

heating up to T¼ 918 K, the coercivity value for the MA

alloy (37.5 A/m) is still lower than that of the BA (43 A/m)

alloy heated above the first crystallization event.17 The

larger coercivity of the devitrified BA can be due to forma-

tion of the Fe23B6-type phase upon annealing, which is

magnetically harder than the (Fe,Co) solid solution precipi-

tated upon annealing of the MA.47 According to the random

anisotropy model,48–50 the exchange interaction will aver-

age out the magnetic anisotropy when the grain size is

smaller than a critical length (L0). In this case, an increase

of crystal size enhances the anisotropy and, consequently,

increases the coercivity. Based on Table III, the coercivity

enhancement with increasing annealing temperature may

be due to the increased magneto-crystalline anisotropy

caused by growth of the (Fe,Co) crystals upon annealing. In

order to prove above statement, the crystallite size of the

devitrified MA alloy was calculated at different annealed

temperatures using Williamson-Hall method.51 The crystal-

lite sizes are 13 6 4 nm, 18 6 4 nm, and 22 6 4 nm at 818 K,

863 K, and 903 K, respectively. It is clear that with increase

in the crystallite the size, the coercivity values also increases

drastically.

The DC hysteresis plots of the MA samples in the

as-cast state and after annealing are shown in Fig. 8. In order

to avoid the differences in the saturation magnetization due

to shape anisotropy, all the samples dimensions were fixed.

For all the measurements, we have used 2 mm diameter sam-

ple with the height of 1 mm. The saturation magnetization of

the MA is slightly higher than for as-cast BA17 (i.e., 109

and 108 Am2/kg, respectively). It can be inferred that the sat-

uration occurs at relatively similar magnetic fields for the

as-cast material and the samples heated up to the end

temperature of the first DSC peak (T¼ 918 K), indicating

that the MA alloy remains soft magnetic upon annealing. In

addition, the saturation magnetization increases after heat

treatment, particularly above T¼ 838 K. The enhanced mag-

netic saturation after annealing up to T¼ 918 K can be attrib-

uted to formation of (Fe,Co) crystals with higher saturation

magnetization than the amorphous phase. According to

Fig. 8, the samples heated up to 993 K exhibit the highest satu-

ration magnetization probably due to formation of other mag-

netic crystals like (Fe,Co)2B and Fe23B6 besides the (Fe,Co)

phase, as shown in Fig. 2. Upon annealing, the BA becomes

more magnetically hard with slight increase in saturation mag-

netization. According to these results, it can be concluded that

the MA exhibits better soft magnetic properties in both as-cast

and annealed conditions compared to the BA.

D. Mechanical and physical properties

Figure 9 shows a typical room temperature compressive

true stress-true strain curves of the 2 mm diameter as-cast

BA (black lines) and MA (red lines) samples. The MA

exhibits a very high yield strength of ry¼ 3523 6 20 MPa,

an elastic strain of ey¼ 1.80%, a very high fracture strength

of rf¼ 3913 6 20 MPa and a considerable plasticity of 0.6%.

In addition, its Young’s modulus, obtained upon linear fitting

the stress-strain curve between 800 MPa and 3200 MPa, is

E¼ 198 6 5 GPa. In contrast to the BA material, the MA

shows a tendency for plastic deformation.11 Shen et al.15

have found that the yielding of bulk samples with similar

chemical composition (i.e., BA with 0.25 at. % Cu) occurs at

3700 MPa, and reported a maximum stress of 4050 MPa and

a plastic strain of 0.6%. They used rod samples with 2 mm

diameter with a composite structure consisting of a-(Fe,Co)

and (Fe,Co)23B6 nanocrystalline grains embedded in an

amorphous matrix. A larger compressive yielding and maxi-

mum stress were reported for the MA glassy samples by

Stoica et al.,35 i.e., 3924 MPa and 4143 MPa, respectively,

together with a larger plastic strain of 1.50%. It is worth to

mention that these data were measured for glassy samples

TABLE III. Coercivity values for the {[(Fe0.5Co0.5)0.75B0.2Si0.05]0.96

Nb0.04}99.5Cu0.5 glass in the as-cast state and after heating to different tem-

peratures. The accuracy of the experimental data lies within 60.1 A/m.

Temperature (K) As-cast 793 813 828 858 888 918

Coercivity Hc (A/m) 2 2 1.5 6.5 10 16 37.5 FIG. 9. Compressive true stress-true strain curve of the {[(Fe0.5Co0.5)0.75

B0.2Si0.05]0.96Nb0.04}99.5Cu0.5 glass measured at a strain rate¼ 1� 10�4 s�1.

073908-7 Ramasamy et al. J. Appl. Phys. 119, 073908 (2016)

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. IP: 147.171.56.188 On: Fri, 19 Feb 2016 16:46:34

Page 9: Kinetic analysis of the non-isothermal crystallization ... · 0.5 bulk metallic glasses were evaluated using differential scanning calorimetry under non-isothermal condition. The

with 1 mm diameter, whereas the current data were measured

for samples with 2 mm diameter. These differences are cer-

tainly generated by the sample geometry. However, it is seen

here that the addition of 0.5% Cu improves the plastic

deformability even for the 2 mm diameter glassy samples.

The MA shows the highest plastic deformability reported for

any other Fe-based BMG with a diameter of 2 mm. In order

to study the effects of the minor crystallization in Cu- con-

taining alloy (MA), the samples were annealed up to 820 K

and 868 K (i.e., Txþ 2 K and Txþ 50 K), respectively, and

subsequently subjected to the compression test. According to

Fig. 10, it can be observed that as soon as the crystals start to

nucleate, no plastic deformation is detectable before fracture

of the annealed samples and they fail prematurely. The sam-

ples heated to 820 K which is slightly higher than Tx (i.e.,

Txþ 2) fails at 3200 6 20 MPa, whereas the sample heated

50 K above the Tx fails at 2250 6 20 MPa. This premature

failure may be due to the presence of nano size crystals

(i.e., at 820 K the crystallite size is approx. 13 6 4 nm) in

the amorphous matrix. The hardness values for the BA and

MA are 1272 and 1242 6 10 HV, respectively. The ratio of

rf/E attains a value of 0.019 and Hv/3E takes a value of

0.02. Hence, the relationship between fracture strength and

hardness rf/E�Hv/3E is verified.52 These results reveal

that the addition of Cu improves the plastic deformability

considerably.

IV. CONCLUSIONS

[(Fe0.5Co0.5)0.75B0.2Si0.05]96Nb4 and {[(Fe0.5Co0.5)0.75

B0.2Si0.05]0.96Nb0.04}99.5Cu0.5 glassy samples were produced

using the injection copper mold technique. Their thermal sta-

bility, crystallization process, and magnetic as well as me-

chanical properties were evaluated. The main conclusions

are as follows:

(1) The addition of 0.5 at. % Cu to the base [(Fe0.5Co0.5)0.75

B0.2Si0.05]96Nb4 alloy promotes the formation of MRO

clusters, which drastically changes the crystallization

behavior. For MA, the crystallization takes place in two

stages: first (Fe,Co) crystals nucleate and in the later

stage (Fe,Co)23B6 crystals form.

(2) The high activation energies for crystallization, Ex

¼ 370 kJ/mol and 340 kJ/mol for BA and MA, respec-

tively, makes these glassy samples resistant against crys-

tallization (diffusion controlled). The activation energy

increases with progress of crystallization, the average

energy required for the completion of the crystallization

process are 645 kJ/mol and 425 kJ/mol for BA and MA,

respectively.

(3) For BA nucleation of crystals takes larger span to com-

plete. In contrast, for MA nucleation occurs immedi-

ately as soon as the material reaches its crystallization

temperature (Tx¼ 818 K), and the crystallization com-

pletes very fast.

(4) The MA remains magnetically very soft even after anneal-

ing, because of the nucleation of (Fe,Co) crystals.

(5) The addition of Cu improves the plastic deformability of

the MA. The combination of a high fracture strength of

3913 MPa and 2.45% fracture strain makes this BMG

attractive for applications.

ACKNOWLEDGMENTS

The work was done in the frame of the EU Initial

Training Network VitriMetTech, FP7-PEOPLE-2013-ITN-

607080. The financial support provided by the “DAAD-IIT

Masters Sandwich Program” is also gratefully acknowledged.

J.E. and M.S. acknowledge the Grant No. ERC-2013-ADG-

340025. The authors would like to thank A. Szabo and A.

B�ardos for Fruitful discussions, as well as B. Bartusch, S.

Donath, and M. Frey for technical assistance.

1A. Inoue, Y. Shinohara, and J. S. Gook, Mater. Trans. JIM 36, 1427

(1995).2T. D. Shen and R. B. Schwarz, Appl. Phys. Lett. 75, 49 (1999).3A. Inoue, A. Takeuchi, and B. Shen, Mater. Trans. 42, 970 (2001).4V. Ponnambalam, S. J. Poon, and G. J. Shiflet, J. Mater. Res. 19, 1320

(2004).5Z. P. Lu, C. T. Liu, J. R. Thompson, and W. D. Porter, Phys. Rev. Lett. 92,

245503 (2004).6C. C. Hays, C. P. Kim, and W. L. Johnson, Phys. Rev. Lett. 84, 2901

(2000).7Y. C. Kim, D. H. Kim, and J.-C. Lee, Mater. Trans. 44, 2224 (2003).8D. H. Bae, M. H. Lee, D. H. Kim, and D. J. Sordelet, Appl. Phys. Lett. 83,

2312 (2003).9A. Inoue and B. L. Shen, Adv. Mater. 16, 2189 (2004).

10P. Gargarella, S. Pauly, M. Samadi Khoshkhoo, U. K€uhn, and J. Eckert,

Acta Mater. 65, 259 (2014).11B. Sarac, B. Klusemann, T. Xiao, and S. Bargmann, Acta Mater. 77, 411

(2014).12X. J. Gu, A. G. McDermott, S. J. Poon, and G. J. Shiflet, Appl. Phys. Lett.

88, 211905 (2006).13A. Inoue, B. L. Shen, and C. T. Chang, Acta Mater. 52, 4093 (2004).14Y. Yoshizawa, S. Oguma, and K. Yamauchi, J. Appl. Phys. 64, 6044

(1988).15B. Shen, H. Men, and A. Inoue, Appl. Phys. Lett. 89, 101915 (2006).16A. Inoue and B. Shen, J. Mater. Res. 18, 2799 (2003).17M. Stoica, R. Li, A. R. Yavari, G. Vaughan, J. Eckert, N. Van Steenberge,

and D. R. Romera, J. Alloys Compd. 504(Suppl. 1), S123 (2010).18M. Avrami, J. Chem. Phys. 7, 1103 (1939).19M. Avrami, J. Chem. Phys. 8, 212 (1940).20M. Avrami, J. Chem. Phys. 9, 177 (1941).21D. Minic and B. Adnad-evic, Thermochim. Acta 474, 41 (2008).

FIG. 10. Compressive true stress-true strain curve of the {[(Fe0.5Co0.5)0.75

B0.2Si0.05]0.96Nb0.04}99.5Cu0.5 glassy samples annealed at 820 K and

868 K, respectively. The measurements were carried out at a strain rate of

1� 10�4 s�1.

073908-8 Ramasamy et al. J. Appl. Phys. 119, 073908 (2016)

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. IP: 147.171.56.188 On: Fri, 19 Feb 2016 16:46:34

Page 10: Kinetic analysis of the non-isothermal crystallization ... · 0.5 bulk metallic glasses were evaluated using differential scanning calorimetry under non-isothermal condition. The

22L. Kong, Y. Gao, T. Song, G. Wang, and Q. Zhai, Thermochim. Acta 522,

166 (2011).23J. M�alek, E. C�erno�skov�a, R. S�vejka, J. S�est�ak, and G. Van der Plaats,

Thermochim. Acta 280–281, 353 (1996).24A. A. Soliman, S. Al-Heniti, A. Al-Hajry, M. Al-Assiri, and G. Al-

Barakati, Thermochim. Acta 413, 57 (2004).25H.-R. Wang, Y.-L. Gao, Y.-F. Ye, G.-H. Min, Y. Chen, and X.-Y. Teng,

J. Alloys Compd. 353, 200 (2003).26J. S. Bl�azquez, C. F. Conde, and A. Conde, Acta Mater. 53, 2305 (2005).27A. Inoue and B. Shen, Mater. Trans. 43, 1230 (2002).28M. Aykol, M. V. Akdeniz, and A. O. Mekhrabov, Intermetallics 19, 1330

(2011).29X. Li, H. Kato, K. Yubuta, A. Makino, and A. Inoue, Mater. Sci. Eng., A

527, 2598 (2010).30H. E. Kissinger, J. Res. Nat. Bur. Stand. 57, 5 (1956).31Y. Jia, S. Zeng, S. Shan, L. Zhang, C. Fan, B. Zhang, Z. Zhan, R. Liu, and

W. Wang, J. Alloys Compd. 440, 113 (2007).32M. Stoica, P. Ramasamy, I. Kaban, S. Scudino, M. Nicoara, G. B. M.

Vaughan, J. Wright, R. Kumar, and J. Eckert, Acta Mater. 95, 335 (2015).33R. Li, S. Kumar, S. Ram, M. Stoica, S. Roth, and J. Eckert, J. Phys. D:

Appl. Phys. 42, 085006 (2009).34M. Ohnuma, K. Hono, H. Onodera, J. S. Pedersen, and S. Linderoth,

Nanostruct. Mater. 12, 693 (1999).35M. Stoica, S. Scudino, J. Bednarcik, I. Kaban, and J. Eckert, J. Appl. Phys.

115, 053520 (2014).

36K. Nakamura, T. Watanabe, K. Katayama, and T. Amano, J. Appl. Polym.

Sci. 16, 1077 (1972).37K. Nakamura, K. Katayama, and T. Amano, J. Appl. Polym. Sci. 17, 1031

(1973).38A. Pratap, K. N. Lad, T. L. S. Rao, P. Majmudar, and N. S. Saxena,

J. Non-Cryst. Solids 345–346, 178 (2004).39A. N. Kolmogorov, Bull. Acad. Sci. USSR, Math. Ser. 1, 355 (1937).40S. Ranganathan and M. Von Heimendahl, J. Mater. Sci. 16, 2401 (1981).41V. R. V. Ramanan and G. E. Fish, J. Appl. Phys. 53, 2273 (1982).42M. E. McHenry, F. Johnson, H. Okumura, T. Ohkubo, V. R. V. Ramanan,

and D. E. Laughlin, Scr. Mater. 48, 881 (2003).43M. E. M. A. Hsiao, D. E. Laughlin, M. J. Kramer, C. Ashe, and T.

Ohkubo, IEEE Trans. Magn. 38, 3039 (2002).44J. Christian, The Theory of Transformations in Metals and Alloys

(Pergamon, Oxford, 2002), p. 529.45R. W. Cahn and P. Haasen, Physical Metallurgy (North Holland, 1996), Vol. 2.46T. Bitoh, A. Makino, and A. Inoue, J. Appl. Phys. 99, 08F102 (2006).47R. S. Sundar and S. C. Deevi, Int. Mater. Rev. 50, 157 (2005).48G. Herzer, Random Anisotropy Model (Springer, Netherlands, 2005), Vol.

184.49G. Herzer, in Handbook of Magnetism and Advanced Magnetic Materials

(John Wiley & Sons, Ltd, 2007).50R. Harris and D. Zobin, J. Phys. F: Met. Phys. 7, 337 (1977).51G. K. Williamson and W. H. Hall, Acta Metall. 1, 22 (1953).52H. S. Chen, Rep. Prog. Phys. 43, 353 (1980).

073908-9 Ramasamy et al. J. Appl. Phys. 119, 073908 (2016)

Reuse of AIP Publishing content is subject to the terms at: https://publishing.aip.org/authors/rights-and-permissions. IP: 147.171.56.188 On: Fri, 19 Feb 2016 16:46:34