You are on page 1of 4

Scripta Materialia 130 (2017) 170173

Contents lists available at ScienceDirect

Scripta Materialia
journal homepage: www.elsevier.com/locate/scriptamat

Regular article

Magnetic entropy change in Gd95Fe2.8Al2.2


amorphous/nanocrystalline ribbons
Qiang Zheng a, Linlin Zhang a,b, Juan Du b,
a

School of Materials and Chemical Engineering, Ningbo University of Technology, Ningbo 315016, People's Republic of China
Key Laboratory of Magnetic Materials and Devices, Ningbo Institute of Material Technology and Engineering, Chinese Academy of Sciences, 1219 Zhongguan West Road, Ningbo 315201,
People's Republic of China

a r t i c l e

i n f o

Article history:
Received 12 May 2016
Received in revised form 5 September 2016
Accepted 29 November 2016
Available online xxxx
Keywords:
Magnetic materials
Amorphous/nanocrystalline materials
Nanocomposites
Magnetocaloric effect

a b s t r a c t
The Gd95Fe2.8Al2.2 amorphous/nanocrystalline composite was obtained by melt spinning. A full width at half
maximum of entropy change SM (FWHM) of 104 K, relatively large refrigerant capacity (RC) of ~551 J kg1
together with SM of 7.8 J kg1 K1 have been achieved at 278 K under a eld change of 5 T. This RC value
is about 35% and 80% larger than those of crystalline Gd (~410 J kg1) and Gd5Si2Ge2 (~306 J kg1) alloy, respectively. The mechanism of the enhanced RC was discussed and would shed a light on searching for excellent magnetic refrigerants at near room temperature.
2016 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Magnetic refrigeration based on magnetocaloric effect (MCE) has


attracted increasing attention because of its superior properties, such
as high energy efciency and environmental friendliness. Therefore, it
is expected to replace traditional vapor-compression refrigeration in
the near future [1]. Currently, the large MCE has been extensively studied on various crystalline materials with a rst-order magnetic transition, such as La (Fe1 x, Six)13 [23], Gd5Si2Ge2 [4], MnFe (P, As) [5],
Ni-Mn-X (X = Sn, In, Sb) [68], etc. The advantage of these crystalline
materials is their large magnetic entropy change. However, due to the
rst-order magnetic transition, the thermal and magnetic hysteresis is
intrinsically unavoidable. Also, the temperature range of full width at
half maximum (FWHM) of the magnetic entropy change (Smax
M ) peak
is only a few to a few tens Kelvin. Therefore, the refrigeration capacity
(RC) in crystalline materials is small.
Amorphous materials undergo second order magnetic transition
(SOMT) which will broaden magnetic entropy change (Sm) peaks
and result in high values of RC [913]. They also have some unique
properties that are superior to those of crystalline alloys, such as, no
thermal and magnetic hysteresis, large electrical resistivity, high
corrosion resistance, tailorable magnetic transition temperature, ne
molding and processing behavior. The negative point of amorphous materials is their lower Curie temperatures and lower magnetic entropy
change compared to those of their crystalline counterparts.
Recently, many efforts have been devoted to improving Curie temperatures and magnetic entropy change of amorphous materials in
Corresponding author.
E-mail address: dujuan@nimte.ac.cn (J. Du).

http://dx.doi.org/10.1016/j.scriptamat.2016.11.041
1359-6462/ 2016 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Gd- systems [1418]. Unfortunately, there are still no amorphous materials having a Curie temperature close to room temperature combined
with a high magnetic entropy change under a magnetic eld change
(5 T). Therefore, it is an open challenge to obtain both high Curie temperature and large magnetic entropy change. In this work, Gd-based
amorphous and nanocrystalline composites are designed based on the
idea of combination effect of amorphous and nanocrystalline in which
the amorphous part contributes to the wide temperature range of
FWHM and the nanocrystalline part gives high magnetic entropy change
and high Tc. In this paper, magnetocaloric effect of Gd95Fe2.8Al2.2
ribbons has been developed. And, a high Sm value of
~7.53 J kg1 K1 at 278 K, RC value of 551 J kg1 and FWHM of 104 K
were obtained under a eld change of 5 T. The mechanism of enhanced
RC was discussed and explained.
Master alloy of Gd95Fe2.8Al2.2 (atomic percent) was prepared by arcmelting a mixture of pure Gd, Fe and Al metals in a titanium-gettered
argon atmosphere. The purities of all raw elements are better than
99.9%. The amorphous ribbons with a thickness of about 20 m and a
width of about 2 mm were produced by melt-spinning technology on
a single-roller copper wheel with a speed of 40 m/s. The structure of
the as-quenched ribbons was examined by X-ray diffraction (XRD)
using Cu-K radiation. The thermal properties were investigated by differential scanning calorimetry (DSC) at a constant heating rate of
40 K/min under puried argon atmosphere. The magnetic properties
were measured by superconducting quantum interference device magnetometer (SQUID).
Fig. 1 shows the XRD patterns of Gd95Fe2.8Al2.2 ribbons. The as-spun
ribbons have sharp Gd peaks imposed on a broad diffused diffraction

Q. Zheng et al. / Scripta Materialia 130 (2017) 170173

171

Fig. 1. X-ray diffraction patterns of as-quenched Gd95Fe2.8Al2.2 ribbons. The inset shows
the DSC curves of Gd95Fe2.8Al2.2 ribbons at a heating rate of 40 K/min.

peak which indicates the amorphous/nanocrystalline composite structure in these ribbons. The amorphous base can be further proved by
the DSC results (the inset of Fig. 1). From the DSC curves, there is a
clear glass transition (Tg) before several exothermic crystallization reactions, which indicates the existing of amorphous material. The sharp Gd
peaks gives the evidence of nanocrystalline exists in the amorphous
base, and the grain size of Gd nanoparticles was approximately 20 nm
calculated by Debye-Scherrer formula.
Fig. 2 presents the temperature dependence of magnetization of
Gd95Fe2.8Al2.2 ribbons under an applied eld of 200 Oe. The Curie temperature (Tc), which is dened as the temperature at the maximum of
dM/dT, where M is magnetization and T is temperature, is shown in
the inset of Fig. 2. There are obviously two maxima of dM/dT curve
which proves the appearance of multiple Curie temperatures. One maximum is located at 232 K, and the other is located at 278 K.
Fig. 3(a) displays the isothermal magnetization curves of
Gd95Fe2.8Al2.2 ribbons measured in a temperature range of 125350 K
under a magnetic eld up to 5 T. From 265 to 290 K, the temperature
step is 5 K and in faraway regions of 125265 K and 290350 K, the temperature step is 20 K. The sweeping rate of the eld was slow enough to
ensure that the magnetization curves are obtained in an isothermal
process. To further understand the magnetic transitions, the Arrot

Fig. 2. Temperature dependence of the magnetization of Gd95Fe2.8Al2.2 ribbons under an


applied eld of 200 Oe. The inset shows the derivative curves of dM/dT vs. T.

Fig. 3. Isothermal magnetization curves between 125 and 350 K (a) and the Arrott plots for
as-quenched Gd95Fe2.8Al2.2 ribbons (b).

plots (M2 versus H/M curves) were also analyzed and shown in Fig.
3(b). According to Banerjee criterion, a magnetic transition is considered as rst-order when the slope of Arrott plot is negative; otherwise,
it is expected to be second-order when the slope is positive [19]. The inection or negative slope as an indication of metamagnetic transition
above Tc is not observed in Fig. 3(b). This proves the magnetic transition
at Tc for the Gd95Fe2.8Al2.2 alloy is a second order magnetic transition
(SOMT) for both the amorphous and nano-Gd parts.
Fig. 4(a) shows the SM-T curve of Gd95Fe2.8Al2.2 ribbons under a
eld change of 5 T. SM for Gd95Fe2.8Al2.2 ribbons was calculated by
using Maxwell relationships. Maximum SM, magnetic transition
temperature (Tc), RC for Gd95Fe2.8Al2.2 ribbons are summarized in
Table 1. For comparison, magnetocaloric properties of some typical
magnetic transition materials around room temperature including our
material are also included. As shown in Fig. 4(a), SM peak of
Gd95Fe2.8Al2.2 ribbon at 283 K are about 2.34 J kg1 K 1 under 1 T,
3.97 J kg1 K1 under 2 T, 5.29 J kg1 K1 under 3 T, 6.45 J kg1 K1
under 4 T and 7.53 J kg1 K1 under 5 T, respectively. For Gd95Fe2.8Al2.2
ribbon, the maximum value of SM under 5 T is comparable to those of
many previously published results, such as Gd5Si2Ge1.9Fe0.1
(7 J kg1 K1, TC = 276 K) [20], La0.7Ca0.2Sr0.1MnO3 (7.45 J kg1 K1,
TC = 308 K) [21]. While, it is much lower than that of other crystalline
materials, such as La1.4Ca1.6Mn2O7 (16.8 J kg1 K1, TC = 270 K) [22],
Gd5Si2Ge2 (18.4 J kg1 K1, TC = 276 K) [4]. However, a large broadening of the SM (T) curves around room temperature can overcompensate for such a disadvantage, giving rise to a signicant RC
enhancement.

172

Q. Zheng et al. / Scripta Materialia 130 (2017) 170173

Fig. 4. The total SM (T) curve (solid square) of Gd95Fe2.8Al2.2 ribbons under a eld
change of 5 T (a). The tting results of the contribution of both amorphous (solid circle)
and nano-Gd parts (solid triangle) to the total SM(T) curve (solid square) are also
included in panel (a), and magnetic eld H dependence of RC and SM of the
Gd95Fe2.8Al2.2 ribbons (b). The dashed line in panel (b) is the tting results of SM Hn.

From Fig. 4(a), it can be seen that the SM(T) curve (solid square)
is asymmetric and could be the result of the convolution of independent
processes for each contribution in the system, which, by the way, is what
can be seen in the thermal magnetization measurement of Fig. 2. The
contributions from both the amorphous (solid circle) and nano-Gd part
(solid triangle) in the composite by tting are displayed in Fig. 4(a).
Since two Tc temperatures have been observed in Fig. 2, two peaks of
SM were intended to obtain by tting SM (T) curve using the function of tting multi-peaks in origin 8.0. The initial two temperature Tc

peaks (two maxima in inset of Fig. 2) were selected for tting the contribution of amorphous and nano-Gd parts. For the tting curves, Gaussian
distribution was applied during the tting temperature range. The coefcient of determination (COD) for nal tting results is 0.992, which means
a good tting result. And for the amorphous part, the SM is
4.0 J kg1 K1at 235 K, RC is 686 J kg1, and FWHM is 230 K. For nanoGd part, the SM is 6.3 J kg1 K1 at 288 K, RC is 331 J kg1, and
FWHM is 72 K. Therefore, the most important point of our research is to
prot the combination effect of amorphous and nanocrystalline phases
on the MCE concerning its temperature range, magnetic entropy variation
and refrigeration capacity. Additionally, according to the characteristic
SM (T) curve of amorphous and crystalline materials, from the tting
results can explain these two maxima in the inset of Fig. 2. The former is
ascribed to the amorphous part, and the latter is due to the precipitated
nano sized Gd.
Also, it can be seen that in the inset of Fig. 4(a), the S of nano-Gd is
about 6.3 J kg1 K1, which is lower than that value (S =
~10 J kg1 K1) in bulk Gd under a magnetic change of 5 T. The reasons
are due to the nite size effect (e.g., a grain size of nanometer size) and
structural disorder at grain boundaries. The above-mentioned reasons
will result in the depressed magnetization (e.g., Magnetic moment)
and then reduced magnetic entropy in nanometer Gd compared to
bulk Gd. Two examples can be found in Ref. [24,25]. In Ref. [24], it was
reported that the delta S was 10.07, 7.73 and 4.47 J kg 1 K1 for
1000 nm, 100 nm and 15 nm Gd systems under a magnetic eld of
5 T, respectively. In Ref. [25], MCE was reported for single crystal Gd,
18 nm and 12 nm nano-Gd under 5 T, and the delta S for nano-Gd
was reduced by roughly a factor of 1.5 compared to that of single crystal
Gd (micrometer).
Another question is why the transition temperature of nano-Gd
lower than that of bulk Gd combined Figs. 2 and 4(a)? Finite size of a
system (e.g., a grain of nanometer size) limits the divergence of and
thereby causes a size-dependent shift in Tc to lower temperatures better
known as the nite-size scaling. For example, in Ref. [25], the Tc is
285.65 K, 287.25 K and 293 K for d = 12 nm, d = 18 nm and single crystal Gd by specic heat measurement.
For the SOMT materials, the relation of SM and H can be described as the following scaling law [26]: S Hn. The tting results
plotted in the Fig. 4(b) demonstrate that the value of n for the meltspun Gd95Fe2.8Al2.2 ribbon is about 0.708. This value is consistent to
the value of 2/3 corresponding to mean eld theory [27].
Refrigeration capacity (RC) is another important parameter characterizing the magnetocaloric effect. In this work, the RC values are determined by numerically integrating the area under the SM (T) curve
with a full width at half maximum of the peak as the integrating limits
[22]. As shown in Fig. 4(a), the FWHM of SM is 104 K and 67 K for a
magnetic eld change of 5 T and 2 T, respectively. And, the RC values
of this composite are about 90 J kg1 under 1 T, 196 J kg1 under 2 T,
309 J kg1 under 3 T, 425 J kg1 under 4 T and 551 J kg1 under 5 T,

Table 1
Maximum magnetic entropy changes (SM), magnetic transition temperature (Tc), RC for our sample together with other magnetocaloric candidate materials at near room temperature
for comparison.
Material

Structure

Applied eld (T)

Transition temperature (K)

SM
(J kg1 K1)

RC (J kg1)

Reference

Gd95Fe2.8Al2.2
Gd50Co45Fe5
Gd48Co52
Gd
Gd5Si2Ge2
Gd5Si2Ge1.9Fe0.1
La0.7Ca0.2Sr0.1MnO3
La1.4Ca1.6Mn2O7
GdFeAl

a+c
a
a
a
c
c
c
c
c

5
5
5
5
5
5
5
5
5

232/278
289
282
294
276
176
308
270
265

7.53
3.8
4.23
10.2
18.6
7
7.45
16.8
3.7

551
521
750
410
306
360

420

This work
[13]
[18]
[1]
[1]
[20]
[21]
[22]
[23]

a and c stand for the amorphous and crystalline structure, respectively.

Q. Zheng et al. / Scripta Materialia 130 (2017) 170173

respectively. The maximum RC value is under 5 T is about 35% and 80%


larger than those of Gd (~410 J kg1) [1] and Gd5Si2Ge2 (~306 J kg1)
[1], respectively.
The Gd95Fe2.8Al2.2 amorphous/nanocrystalline composite was prepared by melt spinning. The Tc is 232 K and 278 K for the amorphous
and nano-Gd parts, respectively. Second-order magnetic transitions
are justied by the Arrott plot for this composite. The large magnetic entropy change of 7.53 J kg 1 K 1 under 5 T was obtained for
Gd95Fe2.8Al2.2 composites. The refrigeration capacity (RC) value is
551 J kg1 under a magnetic eld change of 05 T. The large SM and
high Tc makes Gd-based amorphous/nanocrystalline composites attractive candidate for magnetic refrigeration materials at near room
temperature.
This work has been supported by the Natural Science Foundation of
Zhejiang Province for Outstanding Youth (Grant No. LR12E01001), National Natural Science Foundation of China (Grant No. 51422106), the
Natural Science Foundation of Ningbo City (Grant No. 2015A610004)
and the start-up foundation of Ningbo University of Technology.
References
[1] K.A. Gschneidner Jr., V.K. Pecharsky, A.O. Tsokol, Rep. Prog. Phys. 68 (2005) 1479.
[2] F.X. Hu, B.G. Shen, J.R. Sun, Z.H. Cheng, G.H. Rao, X.X. Zhang, Appl. Phys. Lett. 78
(2001) 3675.
[3] A. Fujita, S. Fujieda, Y. Hasegawa, K. Fukamichi, Phys. Rev. B 67 (2003) 104416.
[4] V.K. Pecharsky, K.A. Gschneidner Jr., Phys. Rev. Lett. 78 (1997) 4494.
[5] O. Tegus, E. Bruck, K.H.J. Buschow, F.R. De Boer, Nature (London) 415 (2002) 150.

173

[6] T. Krenke, E. Duman, M. Acet, E.F. Wassermann, X. Moya, L. Manosa, A. Planes, Nat.
Mater. 4 (2005) 450.
[7] J. Du, Q. Zheng, W.J. Ren, W.J. Feng, X.G. Liu, Z.D. Zhang, J. Phys. D. Appl. Phys. 40
(2007) 5523.
[8] J. Liu, T. Gottschall, K.P. Skokov, J.D. Moore, O. Guteisch, Nat. Mater. 11 (2012) 620.
[9] Q. Luo, D.Q. Zhao, M.X. Pan, R.J. Wang, W.H. Wang, Appl. Phys. Lett. 88 (2006)
181909.
[10] J. Du, Q. Zheng, Y.B. Li, Q. Zhang, D. Li, Z.D. Zhang, J. Appl. Phys. 103 (2008) 023918.
[11] J. Du, Q. Zheng, Y.B. Li, E. Bruck, K.H.J. Buschow, W.B. Cui, W.J. Feng, Z.D. Zhang, J.
Magn. Magn. Mater. 321 (2009) 413.
[12] F. Yuan, J. Du, B.L. Shen, Appl. Phys. Lett. 101 (2012) 032405.
[13] G.L. Liu, D.Q. Zhao, H.Y. Bai, W.H. Wang, M.X. Pan, J. Phys. D. Appl. Phys. 49 (2016)
055004.
[14] Y.T. Wang, H.Y. Bai, M.X. Pan, Sci. China Phys. Mech. Astron. 51 (2008) 337.
[15] B. Schwarz, B. Podmilsak, N. Mattern, J. Eckert, J. Magn. Magn. Mater. 322 (2010)
2298.
[16] Z.G. Zheng, X.C. Zhong, K.P. Su, H.Y. Yu, Z.W. Liu, D.C. Zeng, Sci. China Phys. Mech.
Astron. 54 (2011) 1267.
[17] L. Zhang, M. Bao, Q. Zheng, L. Tian, J. Du, AIP Adv. 6 (2016) 035220.
[18] Z.W. Wang, P. Yu, Y.T. Cui, L. Xia, J. Alloys Compd. 658 (2016) 598.
[19] B.K. Banerjee, Phys. Lett. 12 (1964) 16.
[20] V. Provenzano, A.J. Shapiro, R.D. Shull, Nature 429 (2004) 853.
[21] M.H. Phan, S.C. Yu, N.H. Hur, Appl. Phys. Lett. 86 (2005) 072504.
[22] K.A. Gschneidner Jr., V.K. Pecharsky, A.O. Pecharsky, C.B. Zimm, Mater. Sci. Forum 69
(1999) 315.
[23] Q.Y. Dong, B.G. Shen, J. Chen, J. Shen, H.W. Zhang, J.R. Sun, J. Appl. Phys. 105 (2009),
07A305.
[24] H. Zeng, J. Zhang, C. Kuang, M. Yue, Appl. Nanosci. 1 (2011) 51.
[25] S.P. Mathew, S.N. Kaul, Appl. Phys. Lett. 98 (2011) 172505.
[26] V. Franco, A. Conde, Int. J. Refrig. 33 (2010) 465.
[27] H. Oesterreicher, F.T. Parke, J. Appl. Phys. 55 (1984) 4334.

You might also like