You are on page 1of 8

Journal of Alloys and Compounds 619 (2015) 5865

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


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

Effect of heat treatment on ordering and functional properties


of the Fe19Ga alloy
Ali A. Emdadi a,d,, Joan Cifre b, Olga Yu Dementeva c, Igor S. Golovin d,
a

Faculty of Materials Engineering, Sahand University of Technology, Sahand Town, 51335-1996 Tabriz, Iran
Serveis Cienticotcnics, Universitat de les Illes Balears, Ctra. deValldemossa, km.7.5, E-07122 Palma de Mallorca, Spain
c
Bauman Moscow State Technical University, Baumanskaya str. 2, 5/1, 105005 Moscow, Russia
d
National University of Science and Technology MISIS, Leninsky ave. 4, 119049 Moscow, Russia
b

a r t i c l e

i n f o

Article history:
Received 31 July 2014
Received in revised form 26 August 2014
Accepted 27 August 2014
Available online 16 September 2014
Keywords:
Galfenol
Magnetostrictive behavior
Structural and magnetic ordering
Internal friction

a b s t r a c t
Special attention in this work is paid to the study of the inuence of cooling rates on microstructure and
functional properties, magnetostrictive behavior and internal friction of polycrystalline Fe 19 at.% Ga
samples. After annealing at 900 and 1000 C for 40 min, samples were water quenched (wq) or furnace
cooled (fc). High resolution X-ray diffraction (XRD) exhibits the peak shoulder around 2h = 81 for the
sample furnace-cooled from 900 C (fc900C), which is ascribed to the formation of D03 ordered precipitates. Only a solid solution phase with an A2 structure is recorded in the samples water-quenched from
900 C (wq900C) and 1000 C (wq1000C). Annealing during 2 h at 350 C of the water quenched from
900 C specimen (wqann) also leads to the formation of a shoulder around 2h = 81 due to the appearance
of the D03 phase. Magnetic domain structure studied by magnetic force microscopy (MFM) reveals that
the wq900C sample exhibits a regularly aligned stripe domain structure, while irregular maze-like domain
structure is observed in both the fc900C, and wqann samples. Magnetostriction of the samples is measured
as a function of an externally applied magnetic eld. It was found that the wq1000C sample comprises the
highest magnetostriction, while the fc900C and wqann samples have the lowest magnetostriction against
the applied magnetic eld; consequently, the damping capacity in wq samples is also the highest. Amplitude-, frequency- and temperature dependent internal friction tests of wq, fc, and annealed samples were
carried out. Thermal history dependence of anelastic effects is analyzed on the basis of the A2 ? D03
transformation. Physical mechanisms for observed anelastic effects are discussed. Dilatometry and calorimetry studies are in agreement with the A2 (bcc) ? D03 (bcc-based) ordering phenomenon at 200
300 C in as quenched samples.
2014 Elsevier B.V. All rights reserved.

1. Introduction
FeGa binary alloys, referred to as galfenol, are able to reveal a
mutual relationship between magnetization and strain [1]. This
magnetomechanical coupling feature introduces galfenol alloys
as an appropriate class of smart magnetostrictive materials, especially in the eld of sensor-actuator applications [2]. Unique engineering capabilities of galfenol including high mechanical strength,
acceptable ductility, high damping capacity, machinability, weldability, lower cost, very low hysteresis, limited temperature dependence of magnetomechanical behavior, and easy deposition on a

Corresponding authors at: National University of Science and Technology


MISIS, Leninsky ave. 4, 119049 Moscow, Russia.
E-mail addresses: ali.a.emdadi@gmail.com (A.A. Emdadi), i.golovin@misis.ru
(I.S. Golovin).
http://dx.doi.org/10.1016/j.jallcom.2014.08.231
0925-8388/ 2014 Elsevier B.V. All rights reserved.

silicon substrate offered it as an attractive alternative for peer


Terfenol-D [27].
The FeGa stable and metastable phase diagrams shown in
Fig. 1 suggest several phase transformations depending on cooling
rate. The metastable A2, B2, D03 and modied-D03 (B2-like) phases
as well as the stable L12 may be formed in FeGa binary system
while the Ga content changes from 15% to 30%1 [8]. Structural identication in FeGa alloys by conventional X-ray diffraction (XRD)
methods shows uncertainty, especially at small volume fractions,
in identication of the type of a phase: single phase or phase mixture
[9]. This is due to the identical structures and coherency of A2, B2,
and D03 precipitates and identical X-ray scattering factors between
Fe and Ga [9,10]. The high similarity in the atomic scattering factors
leads to weak super-lattice reections [9].

Only atomic per cents are used in this paper.

A.A. Emdadi et al. / Journal of Alloys and Compounds 619 (2015) 5865

Atomic arrangement highly affects both the value and the sign
of magnetostriction coefcients [11]. Structural studies have indicated that only the B2-like phase plays a positive role for the
enhanced magnetostriction in the FeGa alloys [12], whereas the
D03 phase has a negative effect on magnetostriction [13].
Galfenol alloys have high damping capacity because of their high
or giant magnetostriction. Maximal damping at ADIF curve is proportional to the magnetostriction constant (k): Q1
h.max  k [14]. The
90 domain walls (DWs) are known as the source of high damping
capacity, while 180 DWs practically do not contribute to damping.
Although a few papers have reported on the magnetostrictive
properties, only few papers have been published on internal friction
in this class of alloys, and as yet there are no papers in which amplitude dependent damping and magnetostriction of FeGa are studied
together. This paper is aiming to identify the effect of ordered-D03
precipitates on magnetostriction, amplitude and temperature
dependent anelastic effects, and magnetic domain structures in
Fe19.2Ga, to provide a more comprehensive insight on thermal
history dependency of functional properties in FeGa alloys.
Precipitation of ordered-D03 phase critically affects the magnetostrictive performance as well as damping capacity in FeGa
alloys. D03 precipitates hamper magnetic DWs motion, which
results in a decrease of magnetostriction and damping. Also, the
D03 structure is not conducive to enhanced magnetostriction due
to its negative value of k001 [11]. It is known that water quenching
suppresses the formation of ordered phases and retains the disordered A2 single phase down to room temperature, resulting in
enhanced magnetostriction. In contrast, slow cooling encourages
the precipitating of D03 phase within A2 matrix leading to a phase
mixture of A2 + D03 and degradation the magnetostrictive and
damping properties [9,10,1517].
2. Experimental procedure
Polycrystalline galfenol with a nominal composition Fe80Ga20 was prepared
using commercial pure Fe and 99.999% pure Ga by button-arc melting under the
protection of 99.999% inert argon gas using ALD-LK 645 furnace. After remelting
the button several times to ensure compositional homogeneity, it solidied in a
water-cooled copper mold with a cavity of 30 mm diameter and 10 mm depth. Then
the cast button was homogenized at 900 C for 1 h to eliminate the possible microsegregation. Using energy dispersive spectroscopy (EDS), chemical composition of
the cast button was revealed to have an average composition of Fe19.2 at.% Ga.
To study the phase structure and the structureproperty relation in this alloy,
the samples were cut from the cast button using electro discharge machining
(EDM) and underwent different thermal treatments. The samples were annealed
40 min at 900 and 1000 C in a vacuum tube furnace, then they were cooled in three
paths: in water (wq water quenching), in furnace (fc furnace cooling), and several water quenched samples were additionally annealed at 350 C for 2 h. Consequently, the samples were denoted as wq900C, wq1000C, fc900C, and wqann samples.

59

Phase composition was examined by XRD (Bruker AXS D8 ADVANCE) with Ka1
copper incident radiation of 1.54056 wavelength. In this diffractometer, the beam
is collimated, compressed and frequency ltered by a Gbel mirror and V-Groove to
produce a collimated beam (with a divergence of below 0.007) with dimensions of
about 0.3 mm by 11 mm. The 60 mm Gbel mirror is a multilayer X-ray mirror on a
high precision parabolic surface. The lattice spacing gradient and material combination is optimized for the Cu Ka1 part of the radiation spectrum whereas other
wavelengths are suppressed by several orders of magnitude. This gives a very
reective (95%) optic that not only collimates the beam but acts a monochromator.
The V-Groove is a perfect germanium crystal with a V-shaped groove etched
and polished into it. This compresses the beam via asymmetric (0 2 2) reections
within the Ge crystal. Because it is specic wavelength, it also spectrally cleans
the beam so that only Cu Ka1 is present in the controlled beam.
To increase the resolution of the primary X-ray scans (2h = 20120, step
size = 0.05/s) and to resolve overlapping peaks of different phases, some scans
around 2h  44, 64, 81, 97, and 114 were taken using smaller step size of 0.005/
s. These scans are named as high resolution scan in the text.
Thermal analysis measurements were carried out using Setaram Labsys DSC
1600 equipment between 150 and 700 C in 99.9995 Ar atmosphere with heating
rates from 20 to 40 K/min.
For magnetic force microscopy (MFM) study, 10-mm length specimens were cut
down from heat-treated samples, and then mechanically polished using SiC and alumina suspension as described in Ref. [18]. This polishing procedure minimizes the
residual polishing stress and relatively eliminates the surface damaged layer. A
DualScope C26 MFM was used to attain the magnetic force gradient image with a
typical lift height of 50 nm and a high moment Co-coated tip magnetized normally
to the sample surface. In order to examine the magnetostrictive behavior, a couple of
pin-shaped specimens with the size of 1  5 mm were cut from heat-treated
samples along the diameter direction of cast button. Magnetostrictive strains were
measured in the presence of an externally applied magnetic eld up to saturated
value of 25 kA/m by using a specic experimental setup described in Ref. [19].
Internal friction (IF, or Q1), i.e. tan u at forced vibrations has been measured on
a dynamical mechanical analyser DMA Q800 TA Instruments. The u is the phase lag
between the applied cyclic stress r = r0 cos(xt) and the resulting strain e = e0
cos(xt + u), where x = 2pf and u is the phase or loss angle. These measurements
were conducted as a function of temperature between 0 and 600 C, using forced
bending vibrations in the range between 0.3 and 30 Hz with e0 = 7  105 with a
heating and cooling rate of 1 K/min (temperature dependent IF and elastic modulus: TDIF and TDEM). Additional measurements were conducted as a function of
amplitude between 4  106 and 2  103 at frequency 3 Hz and room temperature
(amplitude dependent internal friction and elastic modulus: ADIF and ADEM,
correspondingly).
Hardness measurements were carried out using universal hardness tester ESEWAY 700 series. The indentation force and the dwell time were set to 1 kgf and 10 s,
respectively. An average value of at least 15 individual measurements was used to
characterize the microhardness.

3. Results
3.1. Phase structure identication
Fig. 2 shows the XRD proles of the samples at different states:
wq900C, wq1000C, fc900C, and wqann. All samples show the (1 1 0),
(2 0 0), (2 1 1), (2 2 0), and (3 1 0) fundamental reections of the

Fig. 1. Stable (a) and metastable (b) FeGa phase diagrams adopted from Ref. [10].

60

A.A. Emdadi et al. / Journal of Alloys and Compounds 619 (2015) 5865
Table 1
Calculated lattice parameters (nm) of the phases in the studied samples.

A2
D03

wq1000C

wq900C

fc900C

wqann

0.29173

0.29169

0.29163
0.58063

0.29068
0.57900

3.2. Magnetic domain structures


Fig. 3 shows magnetic domain patterns of the wq900C, fc900C,
and wqann samples. Stripe-like domains with a high degree of
alignment dominate in the wq900C sample. Average domain width
is about 0.5 lm and their length is above 10 lm. In contrast, irregular maze-like domain patterns with a low degree of alignment are
found in the fc900C and wqann samples.
3.3. Magnetostriction measurements
Fig. 4 shows the magnetostrictive curves of the samples in the
different states. All the measurements were done along the axial
direction of pins, which was strictly parallel to the external eld.
The saturation magnetostriction (kS) of the samples depends on
the cooling rate and it reaches very high values in as quenched
samples. Such an increase in the saturation values of magnetostriction with increase in cooling rate is in agreement with previous
literature [17,20]. The wq1000C sample has the largest magnetostriction value, whereas, the fc900C and wqann samples have the
smallest values.
3.4. Frequency dependent internal friction (FDIF)

Fig. 2. (a) X-ray diffraction spectra for the Fe19.2 at.% Ga samples at different
states. The insets (b) and (c) show high resolution scans of the 81 (2 1 1) reection
showing D03 formed both in the fc900C, and wqannealed samples.

The test conditions for damping measurements were chosen


according to the data shown in Fig. 5, where typical frequency
dependent
curve
(maximal
amplitude
of
deformation
e0 = 7  105) is presented at room temperature for the wq1000C
sample. Dynamical mechanical analyser DMA Q800 TA Instruments sweeps frequency of forced vibrations in the range between
0.01 and 200 Hz. At frequencies above 2030 Hz one can see an
articial peak-like effect due to unfortunate ratio of sample and
apparatus stiffness. At least, from the viewpoint of absolute damping values and scattering of experimental points, damping measurements at frequencies higher than f > 30 Hz (Range II in Fig. 5)
are not reliable. The increase of time for different relaxation processes in the alloy leads to a weak increase of damping with lowering frequency in the Range I. The measurements at frequencies
below 0.3 Hz are too time consuming: one experimental point,
measured by averaging tan u over seven loading cycles, already
at f = 0.1 Hz takes more than 1 min. Usage of low frequencies for
temperature dependent tests with a typical heating and cooling
rate 1 K/min leads to a too low density of experimental points.
Thus, ve frequencies: 0.3, 1, 3, 10 and 30 Hz, marked by arrows
in Fig. 5, have been chosen for our temperature dependent tests
and f = 3 Hz for amplitude dependent tests for a xed temperature.
3.5. Amplitude dependent internal friction (ADIF)

bcc iron with a calculated lattice parameters of 0.29173, 0.29169,


0.29163, and 0.29068 nm, respectively for wq1000C, wq900C,
fc900C, and wqann samples. Insets show the high resolution scans
of the 81 (2 1 1) reection which clearly exhibits the presence of
shoulder peak both for fc900C and wqann samples. Normalization
is done for better comparison between different heat treatment
regimes. The calculated lattice parameters of the phases in the
studied samples are collected in Table 1.

In Fig. 6 one can see ADIF curves for different states of the alloy
(measuring frequency is 3 Hz, test temperature is 30 0.5 C):
water quenched samples from 1000 and 900 C, and annealed at
900 C cooled down in furnace sample. Heat treatment of the Fe
Ga samples inuences the total damping. Water quenched from
1000 C samples were also tested in magnetic eld of attached
magnets: total damping decreases nearly in order of magnitude
in presence of magnets.

A.A. Emdadi et al. / Journal of Alloys and Compounds 619 (2015) 5865

61

Fig. 3. Typical magnetic domain patterns of (a) wq900C, (b) fc900C, and (c) wqann samples.

Fig. 4. Magnetostriction curves for wq1000C, wq900C, fc900C, and wqann specimens.

Fig. 5. Frequency dependent internal friction: f varies from 0.01 to 200 Hz,
e0 = 0.007%, T = 30 C.

The understanding of the complex behavior of the magnetic


characteristics and damping capacity of the FeGa alloy requires
accounting for all the changes in both the ne crystalline structure
and in the magnetic domain structure. Consequently, this has led
us to study the temperature dependent effects.
3.6. Temperature dependent internal friction (TDIF) and elastic
modulus (TDEM)
Fig. 7 shows TDIF and TDEM curves for two states of the Fe
19Ga alloy: the left side gures correspond to water quenched

Fig. 6. Amplitude dependent internal friction curves for wq900C, fc900C, and
wq1000C (without and with magnetic eld) samples. Arrows show inuence: 1
of magnetic eld, 2 of annealing temperature before quenching, and 3 of cooling
rate.

state, and the right side gures to furnace cooled from 900 C
state. Heating and cooling rate of 1 K/min at maximal straine0equals 7  105 and ve chosen frequencies (see Section 3.4) are
used in the temperature range from 0 to 600 C. After heating to
600 C, the samples were cooled down to 0 C under the same measuring conditions. Damping for the wq sample is higher than that
for the fc sample at room temperature. After heating to 600 C
and cooling to room temperature, damping of the wq sample
becomes practically the same as it is in the fc sample. A remarkable
effect takes place in the temperature range from 200 to 300 C: the
elastic modulus E increases with increase in temperature in the wq
sample. A peak-like internal friction effect is observed around
500 C in the studied alloy at heating and cooling cycle. This effect
consists at least of two components: frequency independent and
frequency dependent. Activation parameters for frequency dependent component of the peak can be calculated using frequency
temperature shift of the peak according to Arrhenius equation in
coordinates: ln(1/2pf) vs. 1/T0 [21]. In order to clarify the nature
of the frequency independent effect, the heat ow and dilatometric
tests were carried out.

3.7. Heat ow and dilatometry tests of water quenched samples


Results of the heat ow and the dilatometry tests for wq1000C
sample are shown in Fig. 8. Below 200 C, a normal thermal
expansion takes place. Between 200 and 300 C, there is a relative
volumetric contraction at dilatometry curve, typical for ordering
effect, and an exothermal effect at the DSC curve. In order to
estimate the activation energy of the exothermic peak in the low

62

A.A. Emdadi et al. / Journal of Alloys and Compounds 619 (2015) 5865

Fig. 7. The TDIF and TDEM curves for wq and fc samples at heating (upper raw) to 600 C and subsequent cooling (lower raw): f = 0.3, 1, 3, 10, 30 Hz, e0 = 0.007%.

temperature range (300 C), i.e. the activation energy of the formation of ordered phase, tests with different heating rates were
carried out and treated according to the Kissingers [22] method:

ln b=T 2m Hact =RT m C 2


where b, Tm, and R represents the heating rate, the peak temperature, and the gas constant, respectively. Results are presented in
Fig. 8. The calculated value of activation energy of the D03 ordering
process equals to 0.60 eV, and the Curie temperature was found to
be at 689 C.
A strong expansion effect is observed in 460 C, which according to the phase diagrams may be associated with order
(A2 + D03) ? disorder (A2) transition. According to some existing
data, this temperature is also close to the Curie temperature for
the D03 ordered phase [10]. This effect is recorded at the TDIF

Table 2
Hardness of the specimens in different states.
Sample

wq900C

fc900C

wqann

Hardness Vickers (HV)

250.2 3.3

272.4 4.5

280.5 6.7

curves at 525 C as the frequency independent component of the


internal friction peak (see Fig. 7). The next contraction effect at
630 C on dilatometric curve is due to ferro-to-paramagnetic transition in the A2 phase. The endothermic effect around 700 C on
calorimetric curve conrms this result (the Curie point for Fe
19Ga is 680 C).
3.8. Hardness measurements
Hardness of studied specimens is collected in Table 2. Clearly,
furnace cooling and wq + annealing treatments lead to increase in
hardness value as compared with wq sample. The similar results
have been obtained for the Fe18Ga alloy in Ref. [23] previously.
4. Discussion and conclusions

Fig. 8. Heat ow (20, 30, and 40 K/min) and dilatometry (5 K/min) curves for
wq1000C sample.

Structure and functional properties (high magnetostriction and


damping) of the Fe19Ga alloy were studied after different regimes
of heat treatment, namely (i) different cooling rates (water
quenching and furnace cooling) from single phase range of the
FeGa diagram and (ii) annealing after water quenching. The main
purpose was to nd out and explain correlations between the
structure (appearance of the ordered phases) and the properties.
Very slow phase transformation process of the FeGa alloys
makes the nal structure and their distribution strongly dependent
on heat treatment, which increases the uncertainty of phase

A.A. Emdadi et al. / Journal of Alloys and Compounds 619 (2015) 5865

identication. Existing microstructural studies together with magnetostriction and hardness tests are still incomplete. Moreover, the
full phase identication of the alloys with Ga concentrations more
than 28% has not been reported as yet, and the reason for the sharp
decrease of magnetostriction at higher concentrations of Ga is not
clear enough. Analysis of the results infers that the metastable
phase diagram (Fig. 1a) exhibits better consistence with microstructure-magnetostriction relationships rather than the equilibrium phase diagram (Fig. 1b) [10]. The FeGa metastable phase
diagram (Fig. 1b) suggests that Fe19Ga alloy belongs to the
A2 + D03 two-phase region below 500 C. Thus, the lack of (1 0 0)
and/or (1 1 1) weak super-lattice reection at XRD in several papers
is presumably due to the low volume fraction or very small D03
precipitates [12]. Small D03 domains compared with the X-ray
coherency length will only exhibit the A2 matrix structure [12].
In fact, if the Fe19Ga alloy is quenched rapidly from high temperature single phase down to room temperature, the formation of the
D03 ordered domains is suppressed and the alloy has the single A2
phase structure. On the contrary, when the alloy is cooled down
slowly, the alloy gets the two phases structure A2 + D03; therefore,
the metastable D03 phase may be detached out. Annealing of as
quenched disordered samples should lead to ordering: equilibrium
L12 phase may appear directly from A2 phase or via metastable D03
phase [25,26]. Ordering decreases mobility of magnetic DWs, leading to low magnetostrictive performance [24]. Indeed, the presence
of a shoulder peak at XRD around 2h = 81 for fc900C and wqann
samples indicates the splitting of the (2 1 1) reection and is interpreted as evidence of a two-phase mixture. As discussed in Ref.
[12], the low angle peaks at 80.7945 and 81.0232 represent the
(2 1 1) reection from the A2 phase and the high angle peaks at
81.0716 and 81.3482 represent the (4 2 2) reection from the
D03 ordered phase, respectively. According to Table 1, the calculated lattice parameters of the A2 and D03 phases are 0.29138
and 0.58063 nm respectively for the fc900C and 0.29070 and
0.57900 nm for the wqann samples. Reduction of the A2 lattice
parameter in fc900C and wqann samples (0.29163 and 0.29068 nm,
respectively) in contrast with the wq samples (0.29169 and
0.29173 nm) is the result of the formation of the D03 ordered phase
[16]. Absence of the 81 (2 1 1) peak splitting in the wq samples
suggests the success of quenching treatment in restraining the formation of the ordered phase.
According to magnetostriction measurements, the kS value in
the wq1000C sample is 30 ppm higher than that in the wq900C
sample. This is consistent with previous reports on achieving
higher magnetostriction by using higher quenching temperatures
[25,26]. A reduction in DWs density as well as an augmentation
in homogeneity of DWs motion is caused by both rise in size and
alignment degree of magnetic domains; therefore, improving saturation magnetostriction [27]. As shown in Fig. 3, the stripe-like
domains in the wq900C sample have a much more regular alignment than maze-like domains in fc900C and wqann samples, which
enhances the magnetostriction of the wq900C sample. Moreover,
precipitation of the D03 phase (see Section 3.1) in fc900C and wqann
samples hampers magnetic DWs motion and decreases their
mobility, thus degrading the magnetostrictive response. Quantitatively, the measured value of k001 is negative (107  106 [11])
for the D03 structure. Therefore, the D03 phase structure does not
contribute to enhance magnetostriction in FeGa alloys.
We suppose that the formation of irregular maze-like domain
structures in fc900C and wqann samples (Fig. 3) may be, at least
partly, the result of the precipitation of the ordered phase.
After water quenching from both 1000 and 900 C, the ADIF
curves have pronounced damping peaks due to magnetomechanical damping (Fig. 6). The nature of this effect is magnetomechanical
coupling resulting in damping, i.e. the dissipation of energy of
mechanical vibrations. The magnetomechanical mechanism of this

63

peak is proven by applying an external magnetic eld. In a saturated magnetic eld (MF) magnetic domains inside a sample are
xed and they do not contribute to energy dissipation during vibrations in the elastic range of cyclic loading. The construction of DMA
Q800 apparatus does not allow placing a specimen inside electrical
coil in order to create saturated magnetic eld as it is often used in
torsional pendula. Thus, we attached two NdFeB magnets (Br = 1.2
1.3 kGs, Hci = 1215 kOe, (BH)max = 3035 MGOe) directly to the
sample. Magnets were attached to the centre part of the samples.
Diameter of magnets is 10 mm, working length of the sample
17.5 mm. Thus, magnets cover about 55% of the specimen surface.
This method does not allow xing all magnetic domains in the magnetic eld of attached magnets well, but even this simple experiment demonstrates that total damping becomes signicantly
lower compared to measurements without magnets. Similar inuence of external magnetic eld on damping at torsion was demonstrated earlier for the Fe13Ga [7], Fe18 (Ga + Al) [23] alloys.
There is an important effect in the TDEM and TDIF curves
(Fig. 7a), namely, the elastic modulus E increases in the wq sample
in the temperature range from 200 to 300 C. Increase in modulus
is accompanied by decrease in damping. Similar effect was earlier
reported in Fe20Al alloys at about 130180 C as the result of the
D03 ordering [28]. As it follows from the XRD data, the decrease in
damping after furnace cooling or annealing of quenched samples
is the result of short-range ordering of the sample structures. In
the fc sample these effects are much smaller and they are shifted
to higher temperatures of about 250350 C. Thus, these effects in
quenched sample (increase in modulus and decrease in damping)
are consistent with structural ordering of Ga atoms in Fe and the
decrease in as quenched vacancies. In as-quenched, i.e. disordered
state, FeGa alloys, a structural transition occurs in the range from
200 to 300 C: quenching treatment introduces vacancies and the
short-range D03-type ordering occurs (Ga concentration is not
sufcient for formation of a long range order in this alloy) in
the presence of quenched vacancies [23,29]. This disordering ? short-range ordering structural transformation coincides
with increase of the elastic modulus, as shown in Fig. 7. In the
fc sample density of vacancies and thermodynamical stimulus
are lower than they are in the wq sample, and consequently the
effect of ordering is smaller and it is observed at higher
temperatures.
Fig. 7 shows the peak-like internal friction effect around 500 C
in the studied alloy at heating and cooling cycle. This effect is
different as compared to the similar result reported by one of the
co-authors [7,14,23], where at least two internal friction peaks
were observed in the same temperature range. The relatively
low-temperature thermally-activated internal friction peak at
around 100 C reported in Ref. [7,14] was explained by carbon
atom jumps in the FeGa lattice it is so-called Snoek-type relaxation [30,31]. In contrast with earlier studied alloys, the present
composition does not contain carbon, which is the reason for the
absence of the Snoek-type relaxation peak in the Fe19Ga alloy.
The effect around 500 C in TDIF plot (Fig. 7) is rather complex:
it consists at least of two components. The frequency independent
component of damping is always located at about 500525 C at
heating and 510475 C at cooling. Since it is frequency independent, it may be caused only by structural transition. The second
anelastic component of this peak is frequency dependent.
Overlapping of this frequency dependent component with
frequency independent component makes difculties to evaluate
its activation parameters. Nevertheless, using those peaks, which
maximum is out of the range of frequency independent effect, it
is possible to estimate activation parameters of this thermally
activated relaxation at heating H  2.8 0.4 eV and at cooling
H  2.5 0.3 eV. These values are close to earlier evaluated values
assigned to grain boundary effects in the FeGa alloys [7,14,23].

64

A.A. Emdadi et al. / Journal of Alloys and Compounds 619 (2015) 5865

Heat ow and dilatometry tests give some useful information to


interpret the nature of the frequency independent internal friction
effect around 500 C. Dilatometry plot shows a relative volumetric
contraction between 200 and 300 C and an exothermal effect at
DSC curves which are ascribed to the A2 (bcc) ? D03 (bcc-born)
ordering [32]. In the as-quenched, i.e. in the disordered state
quenching xes non equilibrium vacancies (concentration
qvac  104, time for positron annihilation about 170 ps), thus, a
short-range D03-type ordering phenomenon occurs in the presence
of as-quenched vacancies [23], while Ga concentration is not
sufcient for formation the long-range D03 order in the alloy. The
contraction effect coincides well with the exothermal effect at heat
ow curve: the calculated value of activation energy of the
ordering process by Kissingers method [22] equals to 0.60 eV.
Higher hardness of the fc900C and wq900 + annealed specimens as
compared with wq900 specimen is also the result of structural
ordering in the samples, which decreases mobility of dislocations
and, therefore, increases hardness [23]. These effects: relative
contraction of the sample, increase in elastic modulus, heat ow
peak, and increase in hardness t well to the effect of short-range
ordering at 200300 C, also conrmed by XRD.
Magnetostriction has direct inuence on damping of the sample. The energy loss DW due to magnetic domains for a vibration
stress (r0) below a critical stress (rc) is described by Rayleigh
law: DW = Dr3 (r0 < rc; D = const). If the stresses are large enough
to saturate the energy losses, they can be described by the equation
DW = kkSrc (r0 > rc). Smith and Birchak model states that magnetomechanical damping in ferromagnetic materials originates
from stress-driven irreversible magnetic DWs displacement [33]:
The maximum damping resulting from magnetomechanical effect
(Q1
h.max) is proportional to kSE/ri, where kS, E, and ri correspond
to the saturation magnetostriction, Youngs modulus, and the average internal stress against domain boundary motion, respectively.
The value of maximum hysteretic damping for a Maxwell distribution of internal stresses as: Q1
h.max = 0.34 kkSE/pri(atr0  rmax),
1
where Q1
h.max is the maximum value of Qh in the amplitude dependent damping, k = 1 is a constant characteristic of the shape of the
hysteresis loop. Writing this equation in the form of Q1
h.max = A  kS,
the values of A were calculated from our damping and magnetic
experiments to be 65  106, 50  106 and 45  106, respectively for wq1000C, wq900C, and fc900C.
The small applied periodic deformation (e0 = E1r0) causes the
DWs to carry out small displacements in the proximity of their
equilibrium location: this is the amplitude range in which the
Smith and Birchak model is applicable. The large values of e0 cause
the DWs to move signicantly more, and the inuence of internal
stress is averaged over a signicantly larger region [34]. For such
long-range DWs motion, it is necessary to consider pinning effects
of DWs by different sources including nonmagnetic obstacles, (Krsten theory [35]), magnetic heterogeneities, and the structural heterogeneities including antiphase boundaries etc.
Our results exhibit that the maximum damping value for
wq1000C is maximum (Q1  0.022), for wq900C is moderate
(Q1  0.015), while for fc900C is minimum (Q1  0.011). The Q1
value for wq1000C is about 100% and 47% larger than that for
fc900C and wq900C, respectively. This result is consistent with the
magnetostriction measurements in Fig. 4 showing the maximum
and minimum kS for wq1000C (k = 330 ppm) and fc900C (k = 250
ppm), respectively. In addition, the Q1 and kS values for wq1000C
are about 47% and 10% higher than those for wq900C. It is in agreement with the previous reports demonstrating the larger values of
magnetostriction using higher quenching temperatures [25,26].
The reduction of damping capacity in fc900C sample is due to
the formation of the D03 ordered precipitates during slow cooling,
which was demonstrated by XRD analysis. The same concerns
annealed sample: in this case DSC, dilatometry, IF and E(T) tests

show temperature range of ordering: 200300 C for wq sample


and 250350 C for fc sample (IF, E). The D03 ordering opposes
magnetic DWs displacements and provides additional pinning for
magnetic DWs; therefore, degrading the damping capacity [38].
Water quenching treatment constrains the formation of the D03
structure and preserves the single phase of the A2 structure down
to room temperature; therefore, it leads to enhanced magnetostriction constant and higher damping values. It demonstrates the critical dependence of damping and kS on heat treatment and phase
constituents of the sample. In addition to the A2 phase, which
has positive magnetostriction constant, well-aligned stripe domain
structures also contribute to enhanced magnetostriction and
damping properties of wq sample. Clearly, the more the regularity
of domain structures is, the more the uniformity and simplicity of
magnetic domains rotation in response to external applied magnetic eld. The presence of the D03 ordered domains; acting as
impediments against magnetic DWs motion, along with irregular
domain structures drastically diminishes the magnetostrictive performance and damping capacity of the fc and wqann samples.
Absolute values of damping capacity of studied samples, which
are high but still far from record values [36], depend not only on
the structure of specimens but also on the scheme of loading and
measuring technique. For example, the three point bending at
forced vibrations exhibits slightly higher damping with maximum
of magnetomechanical damping at lower amplitudes as compared
with cantilever tests (experimental results are not shown in this
paper). Increase in static deformation decreases damping and
shifts damping maximum to higher amplitudes. These peculiarities
are the result of different stress distributions in these types of
bending tests (cantilever and three points bending) [37]. Damping
at free decay vibrations is also typically higher than at forced vibrations in cantilever mode.
To sum up:
The XRD analyses show the single phase of the disordered
bcc A2 structure for water-quenched samples, while the
phase mixture of the A2 phase and the structurally-ordered
D03 ordering was detected for both the furnace-cooled and
water-quenched plus annealed samples.
The temperature range of ordering (from 200 to 300 C)
was found in quenched samples by several experimental
methods: DSC, internal friction and modulus, dilatometry
and hardness tests.
The water-quenched samples have mainly well-aligned
stripe domain structure, whereas irregular maze-like
domains were found for furnace-cooled sample. The maximum values of damping and magnetostriction were
achieved for wq1000C, while fc and wqann samples showed
the minimum values. Irregular maze-like magnetic
domains along with the D03 ordered structural domains
lead to weak magnetostrictive response of furnace-cooled
and annealed samples.

Acknowledgements
The reported study was supported by RFBR, Russia, research
project No. 14-03-00165a. Authors are grateful to Prof. D. Hamana
for the dilatometry test, to Dr. S. Hossein Nedjad for the XRD tests,
to Mrs. V. Palacheva for help with DSC tests and data treatment, and
to Mr. M.R. Etminanfar for heat treatment of several specimens.
References
[1] R.A. Kellog, A.B. Flatau, A.E. Clark, M. Wun-Fogle, T.A. Lograsso, J. Appl. Phys. 91
(2002) 7821.

A.A. Emdadi et al. / Journal of Alloys and Compounds 619 (2015) 5865
[2] R.A. Kellogg, Ph.D. Dissertation, Iowa State University, Ames, Iowa, vol. 1, 2003.
[3] R.A. Kellogg, A.M. Russel, T.A. Lograsso, A.B. Flatau, A.E. Clark, M. Wun-Fogle,
Acta Mater. 52 (2004) 5043.
[4] J. Atulasimha, A.B. Flatau, Smt. Mater. Struct. 20 (2011) 043001.
[5] T. Ueno, E. Summers, M. Wun-Fogle, T. Higuchi, Sens. Actuators A 148 (2008)
280.
[6] J.L. Weston, A. Butera, T.A. Lograsso, M. Shamsuzzoha, I. Zana, G. Zangari, J.
Barnard, IEEE Trans. Magn. 38 (2000) 2832.
[7] I.S. Golovin, J. Cifre, J. Alloys Comp. 584 (2014) 322.
[8] O. Ikeda, R. Kainuma, I. Ohnuma, K. Fukamichi, K. Ishida, J. Alloys Comp. 347
(2002) 198.
[9] T.A. Lograsso, A.R. Ross, D.L. Schlagel, A.E. Clark, M. Wun-Fogle, J. Alloys Comp.
350 (2003) 95.
[10] Q. Xing, Y. Du, R.J. McQueeney, T.A. Lograsso, Acta Mater. 56 (2008) 4536.
[11] R. Wu, J. Appl. Phys. 91 (2002) 7358.
[12] T.A. Lograsso, E.M. Summers, Mater. Sci. Eng. A 416 (2006) 240.
[13] Q. Xing, T.A. Lograsso, Scripta Mater. 60 (2009) 373.
[14] I.S. Golovin, A. Rivire, Intermetallics 19 (2011) 453.
[15] N. Srisukhumbowornchai, S. Guruswamy, J. Appl. Phys. 90 (2001) 5680.
[16] J.J. Zhang, T. Ma, M. Van, Phys. B 405 (2010) 3129.
[17] S. Datta, M. Hung, J. Raim, T.A. Lograsso, A.B. Flatau, Mater. Sci. Eng. A 435
(2006) 221.
[18] C. Mudivarthi, S.M. Na, R. Schaefer, M. Laver, M. Wuttig, A.B. Flatau, J. Magn.
Magn. Mater. 322 (2010) 2023.
[19] A.A. Emdadi, S. Hossein Nedjad, H. Badri Ghavifekr, Metall. Mater. Trans. A 45
(2014) 906.

65

[20] H. Yong, D. Yu-tian, Z. Yan-long, W. Guo-bin, Z. Zhi-guang, Trans. Nonferrous


Met. Soc. China 22 (2012) 2146.
[21] M.S Blanter, I.S Golovin, H. Neuhuser, H. Sinning, Handbook of Internal
Friction in Metallic Materials, Springer Verlag, 2007, p. 540.
[22] H.E. Kissinger, Anal. Chem. 29 (1957) 1702.
[23] I.S. Golovin, V.V. Palacheva, V.Yu. Zadorozhnyy, J. Zhu, H. Jiang, J. Cirfe, T.A.
Lograsso, Acta Mater. 78 (2014) 93.
[24] J. Li, X. Gao, Ji. Zhu, J. Li, M. Zhang, J. Alloys Comp. 484 (2009) 203.
[25] A.G. Khachaturyan, D. Viehland, Metall. Mater. Trans. A 38 (2007) 2308.
[26] A.G. Khachaturyan, D. Viehland, Metall. Mater. Trans. A 38 (2007) 2317.
[27] J.X. Zhang, L.Q. Chen, Acta Mater. 53 (2005) 2845.
[28] I.S. Golovin, Anelasticity of iron-based ordered alloys and intermetallic
compounds, in: Y.N. Berdovsky (Ed.), Intermetallics Research Trends, Nova
Science Publishers Inc, 2008, pp. 65133. ISBN 978-1-60021-982-5.
[29] J. Boisse, H. Zapolsky, A.G. Khachaturyan, Acta Mater. 59 (2011) 2656.
[30] I.S. Golovin, S.B. Golovina, Phys. Metals Metallogr. 102 (2006) 593.
[31] S.A. Golovin, I.S. Golovin, Met. Sci. Heat Treat. 54 (2012) 208.
[32] I.S. Golovin, Z. Belamri, D. Hamana, J. Alloys Comp. 509 (2011) 8165.
[33] G.W. Smith, J.R. Birchak, J. Appl. Phys. 39 (1968) 2311.
[34] L. Liao, M. Fang, J. Zhu, J. Li, J. Wang, Int. J. Minerals, Metallurgy Mater. 21
(2014) 1.
[35] V.F. Coronel, D.N. Beshers, J. Appl. Phys. 64 (1988) 2006.
[36] I.S. Golovin, Key Eng. Mater. 319 (2006) 225.
[37] X.F. Hu, S.W. Liu, X.Y. Li, et al., Mater. Sci. Eng. A 528 (2011) 5491.
[38] I.S. Golovin, L. Yu. Dubov, Yu.V. Funtikov et al. Study of ordering and properties
in FeGa alloy with 18 and 21 at% Ga. Submitted to Met. Mat. Trans. A.

You might also like