You are on page 1of 4

Journal of Magnetism and Magnetic Materials 451 (2018) 183–186

Contents lists available at ScienceDirect

Journal of Magnetism and Magnetic Materials


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

Letter to the Editor

Room-temperature ferromagnetism in Fe-based perovskite solid


solution in lead-free ferroelectric Bi0.5Na0.5TiO3 materials
Nguyen The Hung a,b, Luong Huu Bac a, Nguyen Ngoc Trung a,⇑, Nguyen The Hoang a, Pham Van Vinh c,
Dang Duc Dung a,⇑
a
School of Engineering Physics, Ha Noi University of Science and Technology, 1 Dai Co Viet road, Ha Noi, Viet Nam
b
Department of Physics, Faculty of Basic-Fundamental Sciences, Viet Nam Maritime University, 484 Lach Tray road, Ngo Quyen, Hai Phong, Viet Nam
c
Faculty of Physics, Hanoi National University of Education, 136 Xuan Thuy, Cau Giay, Ha Noi, Viet Nam

a r t i c l e i n f o a b s t r a c t

Article history: The integration of ferromagnetism in lead-free ferroelectric materials is important to fabricate smart
Received 28 September 2017 materials for electronic devices. In this work, (1 x)Bi0.5Na0.5TiO3 + xMgFeO3-d materials (x = 0–9 mol%)
Received in revised form 26 October 2017 were prepared through sol–gel method. X-ray diffraction characterization indicated that MgFeO3-d mate-
Accepted 6 November 2017
rials existed as a well solid solution in lead-free ferroelectric Bi0.5Na0.5TiO3 materials. The rhombohedral
Available online 6 November 2017
structure of Bi0.5Na0.5TiO3 materials was distorted due to the random distribution of Mg and Fe cations
into the host lattice. The reduced optical band gap and the induced room-temperature ferromagnetism
Keywords:
were due to the spin splitting of transition metal substitution at the B-site of perovskite Bi0.5Na0.5TiO3
Bi0.5Na0.5TiO3
MgFeO3-d
and the modification by A-site co-substitution. This work elucidates the role of secondary phase as solid
Lead-free ferroelectric solution in Bi0.5Na0.5TiO3 material for development of lead-free multiferroelectric materials.
Multiferroic Ó 2017 Elsevier B.V. All rights reserved.
Sol–gel

1. Introduction indirect exchange interaction through oxygen vacancies, called F


centers [6–10]. However, the magnetization strength of the devel-
Lead-free ferroelectric materials are being rapidly developed to oped materials remains low (as several memu/g) at room temper-
replace lead-based ones because of their environment-friendliness ature; this property hinders the application of such materials for
and safety for human health [1]. real devices. In addition, the paramagnetism of isolated magnetic
Bi-containing perovskite materials have been increasingly ions and/or antiferromagnetic polaron components is strongly
investigated because Bi3+ ions possess lone pairs similar to those affected by the total magnetic moment of the samples. Therefore,
of Pb2+ ions [2,3]. Lead-free ferroelectric Bi0.5Na0.5TiO3-based mate- the magnetization of Bi0.5Na0.5TiO3 materials must be enhanced
rials exhibit piezoelectric and ferroelectric properties that are com- to facilitate their application to electronic devices.
parable with those of Pb(Zr,Ti)O3-based materials [1]. Smolensky The performance of Bi0.5Na0.5TiO3 materials can be improved
et al. fabricated the first Bi0.5Na0.5TiO3 materials in 1960 [4]. These using a solid solution containing various specific perovskite mate-
materials exhibit strong ferroelectric properties, Curie temperature rials. The piezoelectric properties of Bi0.5Na0.5TiO3 materials can be
of 320 °C, remanent polarization of 38 mC/cm2, and coercive field of enhanced by modifying their structure [1]. Guo et al. reported that
73 kV/cm at room temperature [4,5]. Studies also showed that the pyroelectric properties of Bi0.5Na0.5TiO3 were strongly influ-
Bi0.5Na0.5TiO3 materials possess magnetoelectric properties and enced by the addition of Ba(Zr0.055Ti0.945)O3 [11]. Ullah et al.
are thus suitable for fabricating next-generation electronic devices. reported that the electric field-induced strain and dynamic piezo-
In particular, Bi0.5Na0.5TiO3 exerted fact room-temperature ferro- electric coefficient of Bi0.5Na0.5TiO3 materials were enhanced due
magnetism as a result of the substitution of transition metals, such to modification of BiAlO3 materials [12]. However, Kim et al.
as Fe, Co, Mn, and Cr, into the octahedral Ti site [6–9]. The room- reported that doping LiNbO3 into Bi0.5Na0.5TiO3 affected the dielec-
temperature ferromagnetism in ferroelectric Bi0.5Na0.5TiO3 materi- tric behavior and the ferroelectric properties of the latter; that is,
als doped with transition metals possibly originates from the the reduction in the depolarization temperature corresponds to
the conversion of the ferroelectric phase into the paraelectric phase
[13].
⇑ Corresponding author. In this work, strong room-temperature ferromagnetism in
E-mail addresses: trung.nguyenngoc@hust.edu.vn (N.N. Trung), dung.dangduc Bi0.5Na0.5TiO3 materials was obtained by modifying MgFeO3-d as
@hust.edu.vn (D.D. Dung).

https://doi.org/10.1016/j.jmmm.2017.11.015
0304-8853/Ó 2017 Elsevier B.V. All rights reserved.
184 N.T. Hung et al. / Journal of Magnetism and Magnetic Materials 451 (2018) 183–186

solid solution. The optical band gap of Bi0.5Na0.5TiO3 materials coordination number 8) and 0.72 Å (in coordination number 6);
decreased from 3.09 eV for pure Bi0.5Na0.5TiO3 to 2.43 eV for 9 moreover, of the radius of Fe2+/3+ (0.770 Å/0.645 Å in coordination
mol% MgFeO3-d-added Bi0.5Na0.5TiO3 materials. The induced number 6) is larger than that of Bi3+ (1.17 Å in coordination num-
room-temperature ferromagnetism and the reduced optical band ber 12), Na+ (1.39 Å in coordination number 12), and Ti4+ (0.605 Å
gap of Bi0.5Na0.5TiO3 materials were due to the random distribution in coordination number 6), respectively [16]. The expansion of the
of Mg and Fe cations into the host lattice. lattice parameters could be due to the substitution of Mg and Fe
cations at the Ti-site rather than at the Bi/Na-site. In addition,
the impurity phase was not detected under the resolution of XRD
2. Experimental
method. The randomly distributed Mg and Fe cations in the crystal
lattice of Bi0.5Na0.5TiO3 altered the phonon vibration (Fig. 1(c)).
(1-x)Bi0.5Na0.5TiO3 + xMgFeO3-d solid solution materials (range
The Raman scattering of undoped and MgFeO3-d-modified
x = 0–9 mol% (denoted as BNT-xMgFeO3-d)) were prepared through
Bi0.5Na0.5TiO3 materials showed a broad band, which could be sep-
a sol–gel technique [8,9]. The composition of the samples was
arated into three region modes due to the random distribution of Bi
determined by electron probe microanalysis. The crystalline struc-
and Na cations to the A-site. The first bands are mainly due to the
tures and vibration modes of the samples were determined
vibration of B-site cations, i.e., Ti–O bond, and the vibration mode
through X-ray diffraction (XRD) and Raman spectroscopy analyses,
is assigned to A1(TO) [17]. The secondary bands are related to the
respectively. Optical properties were studied by UV–Vis spec-
vibration of the TiO6 octahedra and assigned to A1(LO), A1(TO), and
troscopy. Magnetic properties were characterized by a vibrating
E(LO) modes [18,19]. The last region bands correspond to the
sample magnetometer at room temperature.
rhombohedral lattice containing the octahedral distorted [TiO6]
clusters [20]. The overlapping of the Raman peaks was distin-
3. Results and discussion guished using Lorentzian fitting for the undoped and Bi0.5Na0.5TiO3
materials modified with 3 and 5 mol% MgFeO3-d (Fig. 1(d)). The
Fig. 1(a) shows the XRD pattern of undoped and MgFeO3-d- eight phonon vibration modes were recorded and indexed, consis-
modified Bi0.5Na0.5TiO3 materials with various MgFeO3-d concen- tent with the calculation for Raman modes of Bi0.5Na0.5TiO3 mate-
trations. The peak positions and relative intensity match well with rials [17]. The modes tended to shift to low frequencies; the
the perovskite structure and possess the rhombohedral symmetry, differences in the mass of Mg and Fe cations from that of the host
indicating that MgFeO3-d materials existed as a well solid solution Ti and the type of binding with oxygen both changed the frequency
in Bi0.5Na0.5TiO3 materials. The results were validated by applying modes [8,20]. These results provide a solid evidence for the
the Hume–Rothery rules considering the radius of the cations random substitution of Mg and Fe cations into the lattice of
[14,15]. Fig. 1(b) shows that Mg and Fe cations distorted the struc- Bi0.5Na0.5TiO3 materials. Hence, MgFeO3-d materials existed as a
ture of Bi0.5Na0.5TiO3 materials; in the image, the XRD patterns of well solid solution in Bi0.5Na0.5TiO3 materials.
setline (0 1 2)/(1 1 0) peaks were magnified from 31° to 34°. The Fig. 2(a) shows the absorbance spectra of undoped and
peaks tended to shift to low angles, thereby expanding the lattice MgFeO3-d-added Bi0.5Na0.5TiO3 materials. MgFeO3-d-added
parameter. Shannon reported that the radii of Mg2+ are 0.89 Å (in Bi0.5Na0.5TiO3 samples red shifted the absorption edge. The random

Fig. 1. (a) X-ray diffraction pattern of MgFeO3-d solid solution into Bi0.5Na0.5TiO3 with various concentrations within the 2h range of 20–70°; (b) Magnified X-ray diffraction
within 2h range of 31–34° for comparing setline (0 1 2)/(1 1 0) peaks; (c) Raman scattering spectra of MgFeO3-d solid solution into Bi0.5Na0.5TiO3 with various concentrations
at 200–1000 cm 1; and (d) deconvolution Raman peaks of pure Bi0.5Na0.5TiO3, Bi0.5Na0.5TiO3-3MgFeO3-d. and Bi0.5Na0.5TiO3-5MgFeO3-d.
N.T. Hung et al. / Journal of Magnetism and Magnetic Materials 451 (2018) 183–186 185

Fig. 2. (a) UV-Vis absorption spectra of MgFeO3-d-modified Bi0.5Na0.5TiO3 samples as a function of MgFeO3-d concentration; and (b) the (ahm)2 proposal with photon energy
(hm) of Bi0.5Na0.5TiO3 samples as a function of the amount of MgFeO3-d added. The inset of (b) shows the optical band gap Eg value of Bi0.5Na0.5TiO3 samples as a function of the
amount of MgFeO3-d added.

distribution of Mg and Fe cations in the crystal structure of Bi0.5- undoped Bi0.5Na0.5TiO3 samples possessed anti-S shape, which
Na0.5TiO3 led to changes in the electronic band structure. In addi- could contribute to diamagnetic and weak ferromagnetic proper-
tion, the appearance of a tail around 487 nm in the absorbance ties. Clear hysteresis loops were obtained in the undoped Bi0.5-
band could be related to the local transition of Fe cations due to Na0.5TiO3 samples after subtracting the diamagnetic component
spin splitting under the crystal field [21]. Optical band gap energy (inset of Fig. 3). The diamagnetism and weak ferromagnetism of
(Eg) was estimated using linear fitting from photon energy (hm) the undoped Bi0.5Na0.5TiO3 samples possibly originated from the
dependent (ahm)2 (Fig. 2(b)). The Eg values were plotted as a func- empty 3d shell of Ti4+ and Ti4+ or O vacancies [8,9,25]. The M-H
tion of MgFeO3-d amount, as shown in the inset of Fig. 2(b); the val- curves showed the S-shape of the material added with MgFeO3-d
ues decreased from 3.09 eV for undoped to 2.43 eV for 9 mol% and the saturation of the magnetization in the material added with
MgFeO3-d-added Bi0.5Na0.5TiO3 samples. The reduction in the opti- lower than 3 mol% MgFeO3-d; the material added with high con-
cal band gap in Bi0.5Na0.5TiO3 materials could be attributed to the centrations of MgFeO3-d showed the unsaturation of magnetization
following: i) spin slitting of transition metals in the crystal field, upon the application of the magnetic field because of the contribu-
ii) oxygen vacancies due to unbalanced charge between dopants tion of the paramagnetism of the isolated Fe cations and/or antifer-
(i.e., Fe2+/3+, Mg2+) and Ti4+, iii) changes in the bonding type romagnetism of the magnetic polaron interaction [6–10]. The non-
between hybridization A–O in general ABO3 perovskite, and iv) sur- zero remanence and coercivity obtained in the M-H curves were
face effect on nanocrystal size [8,9,21–24]. Thus, the reduction in solid evidence of the ferromagnetic ordering at room temperature.
the optical band gap could improve the photovoltaic and photocat- The magnetization strength in MgFeO3-d-added Bi0.5Na0.5TiO3
alytic performances of ferroelectric perovskite materials. materials was enhanced to around 39.6 memu/g (around 18.7
MgFeO3-d-modified Bi0.5Na0.5TiO3 samples exhibited enhanced memu/g after subtracting the paramagnetic-like component) at 6
ferromagnetism at room temperature. Magnetization was depen- kOe. This value is higher than that of Bi0.5Na0.5TiO3 materials doped
dent on the applied strength in the magnetic field (M-H) curves with single Mn (9 memu/g), Cr (1.5 memu/g), Fe (11 memu/
of undoped and MgFeO3-d-added Bi0.5Na0.5TiO3 samples with vari- g), and Co (3 memu/g) [6–9]. Coey et al. pointed out that the
ous MgFeO3-d concentrations at room temperature (Fig. 3). The interaction of magnetic ions through oxygen vacancies was
strongly dependent on the hydrogenic orbital of the effective
radius [10,26]. Thus, we suggest that the ferromagnetic interaction
was enhanced because Mg cations controlled the distance and
interaction of magnetic ions through oxygen vacancies.

4. Conclusion

MgFeO3-d materials existed as a well solid solution in Bi0.5Na0.5-


TiO3 materials. The random distribution of Mg and Fe cations into
the lattice of Bi0.5Na0.5TiO3 reduced the optical band gap and
strongly influenced room-temperature ferromagnetism. This work
provides a basis for integrating ferromagnetism in lead-free ferro-
electric materials to fabricate green electronic devices.

Acknowledgments

Fig. 3. M-H curves of the pure Bi0.5Na0.5TiO3 and MgFeO3-d-modified Bi0.5Na0.5TiO3 This work was financially supported by The Ministry of Educa-
samples added with various amounts of MgFeO3-d at room temperature. tion and Training, Viet Nam, under project number B2016-BKA-25.
186 N.T. Hung et al. / Journal of Magnetism and Magnetic Materials 451 (2018) 183–186

References [13] J.S. Kim, C.H. Chung, H.S. Lee, S.T. Chung, J. Korean Phys. Soc. 58 (2011) 659–
662.
[14] W. Hume-Rothery, Atomic Theory for Students of Metallurgy, The Institute of
[1] N.D. Quan, L.H. Bac, D.V. Thiet, V.N. Hung, D.D. Dung, Adv. Mater. Sci. Eng. 2014
Metals, London, 1969 (fifth reprint).
(2014), article ID 365391.
[15] C. Barry Carter, M. Grant Norton, Ceramic Materials: Science and Engineering,
[2] P. Baettig, C.F. Schelle, R. Lesar, U.V. Waghmare, N.A. Spaldin, Chem. Mater. 17
Springer, 2007, pp. 126.
(2005) 1376–1380.
[16] R.D. Shannon, Acta. Cryst. A 32 (1976) 751–767.
[3] X. He, K.J. Jin, Phys. Rev. B 94 (2016) 224107.
[17] M.K. Niranjan, T. Karthik, S. Asthana, J. Pan, J. Appl. Phys. 113 (2013) 194106.
[4] G.A. Smolensky, V.A. Isupov, A.I. Agranovskaya, N.N. Krainic, Fizika Tverdogo
[18] J. Kreisel, A.M. Glazer, P. Bouvier, G. Lucazeau, Phys. Rev. B 63 (2001) 174106.
Tela. 2 (1960) 2982–2985.
[19] Y. Chen, K.H. Lam, D. Zhou, J.Y. Dai, H.S. Luo, X.P. Jiang, H.L.W. Chan, Inter.
[5] T. Takenaka, K.I. Maruyama, K. Sakata, Jpn. J. Appl. Phys. 30 (1991) 2236–2239.
Ferroelectric 141 (2013) 120–127.
[6] Y. Wang, G. Xu, L. Yang, Z. Ren, X. Wei, W. Weng, P. Du, G. Shen, G. Shen, G. Han,
[20] J. Suchanicz, I.J. Sumara, T.V. Kruzina, J. Electroceram. 27 (2011) 45–50.
Mater. Sci. Poland 27 (2009) 471–476.
[21] D.D. Dung, D.V. Thiet, D. Odkhuu, L.V. Cuong, N.H. Tuan, S. Cho, Mater. Lett. 156
[7] Y. Wang, G. Xu, X. Ji, Z. Ren, W. Weng, P. Du, G. Shen, G. Han, J. Alloys
(2015) 129–133.
Compound. 475 (2009) L25–L30.
[22] N.D. Quan, V.N. Hung, N.V. Quyet, H.V. Chung, D.D. Dung, AIP Advances 4
[8] L.T.H. Thanh, N.B. Doan, L.H. Bac, D.V. Thiet, S. Cho, P.Q. Bao, D.D. Dung, Mater.
(2014) 017122.
Lett. 186 (2017) 239–242.
[23] N.V. Quyet, L.H. Bac, D. Odkhuu, D.D. Dung, J. Phys. Chem. Solids 85 (2015)
[9] L.T.H. Thanh, N.B. Doan, N.Q. Dung, L.V. Cuong, L.H. Bac, N.A. Duc, P.Q. Bao, D.D.
148–154.
Dung, J. Electron. Mater. 46 (2017) 3367–3372.
[24] L.H. Bac, L.T.H. Thanh, N.V. Chinh, N.T. Khoa, D.V. Thiet, T.V. Trung, D.D. Dung,
[10] J.M.D. Coey, M. Venkatesan, C.B. Fitzgerals, Nature Mater. 4 (2005) 173–179.
Mater. Lett. 164 (2016) 631–635.
[11] F. Guo, B. Yang, S. Zhang, F. Wu, D. Liu, P. Hu, Y. Sun, D. Wang, W. Cao, Appl.
[25] Y. Zhang, J. Hu, F. Gao, H. Liu, H. Qin, Comput, Theor. Chem. 967 (2011) 284–
Phys. Lett. 103 (2013) 182906.
288.
[12] A. Ullah, C.W. Ahn, K.B. Jang, A. Hussain, I.W. Kim, Ferroelectrics 404 (2010)
[26] K. Balamurugan, N.H. Kumar, J.A. Chelvane, P.N. Santhosh, Physica B 407
167–172.
(2012) 2519–2523.

You might also like