Anda di halaman 1dari 16

Accepted Manuscript

Letter to the Editor

Room-temperature ferromagnetism in Fe-based perovskite solid solution in


lead-free ferroelectric Bi0.5Na0.5TiO3 materials

Nguyen The Hung, Luong Huu Bac, Nguyen Ngoc Trung, Nguyen The Hoang,
Pham Van Vinh, Dang Duc Dung

PII: S0304-8853(17)33073-1
DOI: https://doi.org/10.1016/j.jmmm.2017.11.015
Reference: MAGMA 63355

To appear in: Journal of Magnetism and Magnetic Materials

Received Date: 28 September 2017


Revised Date: 26 October 2017
Accepted Date: 6 November 2017

Please cite this article as: N.T. Hung, L.H. Bac, N.N. Trung, N.T. Hoang, P. Van Vinh, D.D. Dung, Room-
temperature ferromagnetism in Fe-based perovskite solid solution in lead-free ferroelectric Bi0.5Na0.5TiO3 materials,
Journal of Magnetism and Magnetic Materials (2017), doi: https://doi.org/10.1016/j.jmmm.2017.11.015

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Room-temperature ferromagnetism in Fe-based perovskite solid solution in
lead-free ferroelectric Bi0.5Na0.5TiO3 materials

Nguyen The Hung1,2, Luong Huu Bac1, Nguyen Ngoc Trung1, Nguyen The Hoang1,Pham Van
Vinh3, and Dang Duc Dung1,*
1
School of Engineering Physics, Ha Noi University of Science and Technology, 1 Dai Co Viet
road, Ha Noi, Viet Nam

2
Department of Physics, Faculty of Basic-Fundamental Sciences, Viet Nam Maritime
University, 484 Lach Tray road, Ngo Quyen, Hai Phong, Viet Nam

3
Faculty of Physics, Hanoi National University of Education, 136 Xuan Thuy, Cau Giay, Ha
Noi, Viet Nam

Abstract
The integration of ferromagnetism in lead-free ferroelectric materials is important to fabricate
smart materials for electronic devices. In this work, (1-x)Bi0.5Na0.5TiO3 + xMgFeO3- materials
(x = 09 mol%) were prepared through solgel method. X-ray diffraction characterization
indicated that MgFeO3- materials existed as a well solid solution in lead-free ferroelectric
Bi0.5Na0.5TiO3 materials. The rhombohedral 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 were due to the spin splitting of
transition metal substitution at the B-site of perovskite Bi0.5Na0.5TiO3 and the modification by A-
site co-substitution. This work elucidates the role of secondary phase as solid solution in
Bi0.5Na0.5TiO3 material for development of lead-free multiferroelectric materials.

Keywords: Bi0.5Na0.5TiO3; MgFeO3-; lead-free ferroelectric; multiferroic; sol-gel.

*
) Corresponding e-mail: dung.dangduc@hust.edu.vn

1
I. Introduction

Lead-free ferroelectric materials are being rapidly developed to replace lead-based ones

because of their environment-friendliness and safety for human health [1]. Bi-containing

perovskite materials have been increasingly investigated because Bi3+ ions possess lone pairs

similar to those of Pb2+ ions [2, 3]. Lead-free ferroelectric Bi0.5Na0.5TiO3-based materials exhibit

piezoelectric and ferroelectric properties that are comparable with those of Pb(Zr,Ti)O3-based

materials [1]. Smolensky et al. fabricated the first Bi0.5Na0.5TiO3 materials in 1960 [4]. These

materials exhibit strong ferroelectric properties, Curie temperature of 320 C, remanent

polarization of 38 C/cm2, and coercive field of 73 kV/cm at room temperature [4, 5]. Studies

also showed that Bi0.5Na0.5TiO3 materials possess magnetoelectric properties and are thus

suitable for fabricating next-generation electronic devices. In particular, Bi0.5Na0.5TiO3 exerted

fact room-temperature ferromagnetism as a result of the substitution of transition metals, such

as Fe, Co, Mn, and Cr, into the octahedral Ti site [69]. The room-temperature ferromagnetism

in ferroelectric Bi0.5Na0.5TiO3 materials doped with transition metals possibly originates from

the indirect exchange interaction through oxygen vacancies, called F centers [610]. However,

the magnetization strength of the developed materials remains low (as several memu/g) at room

temperature; this property hinders the application of such materials for real devices. In addition,

the paramagnetism of isolated magnetic ions and/or antiferromagnetic polaron components is

strongly affected by the total magnetic moment of the samples. Therefore, the magnetization of

Bi0.5Na0.5TiO3 materials must be enhanced to facilitate their application to electronic devices.

The performance of Bi0.5Na0.5TiO3 materials can be improved using a solid solution containing

various specific perovskite materials. The piezoelectric properties of Bi0.5Na0.5TiO3 materials

2
can be enhanced by modifying their structure [1]. Guo et al. reported that the pyroelectric

properties of Bi0.5Na0.5TiO3 were strongly influenced by the addition of Ba(Zr0.055Ti0.945)O3

[11]. Ullah et al. reported that the electric field-induced strain and dynamic piezoelectric

coefficient of Bi0.5Na0.5TiO3 materials were enhanced due to modification of BiAlO3 materials

[12]. However, Kim et al. reported that doping LiNbO3 into Bi0.5Na0.5TiO3 affected the

dielectric behavior and the ferroelectric properties of the latter; that is, the reduction in the

depolarization temperature corresponds to the conversion of the ferroelectric phase into the

paraelectric phase [13].

In this work, strong room-temperature ferromagnetism in Bi0.5Na0.5TiO3 materials was obtained

by modifying MgFeO3- as solid solution. The optical band gap of Bi0.5Na0.5TiO3 materials

decreased from 3.09 eV for pure Bi0.5Na0.5TiO3 to 2.43 eV for 9 mol% MgFeO3--added

Bi0.5Na0.5TiO3 materials. The induced room-temperature ferromagnetism and the reduced

optical band gap of Bi0.5Na0.5TiO3 materials were due to the random distribution of Mg and Fe

cations into the host lattice.

II. Experimental

(1-x)Bi0.5Na0.5TiO3 + xMgFeO3- solid solution materials (range x = 09 mol% (denoted as

BNT-xMgFeO3-)) were prepared through a solgel technique [8, 9]. The composition of the

samples was determined by electron probe microanalysis. The crystalline structures and

vibration modes of the samples were determined through X-ray diffraction (XRD) and Raman

spectroscopy analyses, respectively. Optical properties were studied by UVVis spectroscopy.

Magnetic properties were characterized by a vibrating sample magnetometer at room

temperature.

3
III. Results and discussion

Fig. 1 (a) shows the XRD pattern of undoped and MgFeO3--modified Bi0.5Na0.5TiO3 materials

with various MgFeO3- concentrations. The peak positions and relative intensity match well with

the perovskite structure and possess the rhombohedral symmetry, indicating that MgFeO3-

materials existed as a well solid solution in Bi0.5Na0.5TiO3 materials. The results were validated

by applying the HumeRothery rules considering the radius of the cations [14, 15]. Fig. 1(b)

shows that Mg and Fe cations distorted the structure of Bi0.5Na0.5TiO3 materials; in the image,

the XRD patterns of setline (012)/(110) peaks were magnified from 31 to 34. The peaks

tended to shift to low angles, thereby expanding the lattice parameter. Shannon reported that the

radii of Mg2+ are 0.89 (in coordination number 8) and 0.72 (in coordination number 6);

moreover, of the radius of Fe2+/3+ (0.770 /0.645 in coordination number 6) is larger than that

of Bi3+ (1.17 in coordination number 12), Na+ (1.39 in coordination number 12), and Ti4+

(0.605 in coordination number 6), respectively [16]. The expansion of the 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 method. The

randomly distributed Mg and Fe cations in the crystal lattice of Bi0.5Na0.5TiO3 altered the

phonon vibration [Fig. 1(c)]. The Raman scattering of undoped and MgFeO3--modified

Bi0.5Na0.5TiO3 materials showed a broad band, which could be separated into three region

modes due to the random distribution of Bi and Na cations to the A-site. The first bands are

mainly due to the vibration of B-site cations, i.e., TiO bond, and the vibration mode is assigned

to A1(TO) [17]. The secondary bands are related to the vibration of the TiO6 octahedra and

assigned to A1(LO), A1(TO), and E(LO) modes [18, 19]. The last region bands correspond to

the rhombohedral lattice containing the octahedral distorted [TiO6] clusters [20]. The

4
overlapping of the Raman peaks was distinguished using Lorentzian fitting for the undoped and

Bi0.5Na0.5TiO3 materials modified with 3 and 5 mol% MgFeO3- [Fig.1 (d)]. The eight phonon

vibration modes were recorded and indexed, consistent with the calculation for Raman modes of

Bi0.5Na0.5TiO3 materials [17]. The modes tended to shift to low frequencies; the differences in

the mass of Mg and Fe cations from that of the host Ti and the type of binding with oxygen both

changed the frequency modes [8, 20]. These results provide a solid evidence for the random

substitution of Mg and Fe cations into the lattice of Bi0.5Na0.5TiO3 materials. Hence, MgFeO3-

materials existed as a well solid solution in Bi0.5Na0.5TiO3 materials.

Fig. 2(a) shows the absorbance spectra of undoped and MgFeO3--added Bi0.5Na0.5TiO3

materials. MgFeO3--added Bi0.5Na0.5TiO3 samples red shifted the absorption edge. The random

distribution of Mg and Fe cations in the crystal structure of Bi0.5Na0.5TiO3 led to changes in the

electronic band structure. In addition, the appearance of a tail around 487 nm in the absorbance

band could be related to the local transition of Fe cations due to spin splitting under the crystal

field [21]. Optical band gap energy (Eg) was estimated using linear fitting from photon energy

(h) dependent (h)2 [Fig. 2(b)]. The Eg values were plotted as a function of MgFeO3-

amount, as shown in the inset of Fig. 2(b); the values decreased from 3.09 eV for undoped to

2.43 eV for 9 mol% MgFeO3--added Bi0.5Na0.5TiO3 samples. The reduction in the optical band

gap in Bi0.5Na0.5TiO3 materials could be attributed to the following: i) spin slitting of transition

metals in the crystal field, ii) oxygen vacancies due to unbalanced charge between dopants (i.e.,

Fe2+/3+, Mg2+) and Ti4+, iii) changes in the bonding type between hybridization AO in general

ABO3 perovskite, and iv) surface effect on nanocrystal size [8, 9, 2124]. Thus, the reduction in

the optical band gap could improve the photovoltaic and photocatalytic performances of

ferroelectric perovskite materials.

5
MgFeO3--modified Bi0.5Na0.5TiO3 samples exhibited enhanced ferromagnetism at room

temperature. Magnetization was dependent on the applied strength in the magnetic field (M-H)

curves of undoped and MgFeO3--added Bi0.5 Na0.5TiO3 samples with various MgFeO3-

concentrations at room temperature (Fig. 3). The undoped Bi0.5Na0.5TiO3 samples possessed

anti-S shape, which could contribute to diamagnetic and weak ferromagnetic properties. Clear

hysteresis loops were obtained in the undoped Bi0.5Na0.5TiO3 samples after subtracting the

diamagnetic component (inset of Fig. 3). The diamagnetism and weak ferromagnetism of the

undoped Bi0.5Na0.5TiO3 samples possibly originated from the empty 3d shell of Ti4+ and Ti4+ or

O vacancies [8, 9, 25]. The M-H curves showed the S-shape of the material added with MgFeO3-

and the saturation of the magnetization in the material added with lower than 3 mol%

MgFeO3-; the material added with high concentrations of MgFeO3- showed the unsaturation of

magnetization upon the application of the magnetic field because of the contribution of the

paramagnetism of the isolated Fe cations and/or antiferromagnetism of the magnetic polaron

interaction [6-10]. The non-zero remanence and coercivity obtained in the M-H curves were

solid evidence of the ferromagnetic ordering at room temperature. The magnetization strength in

MgFeO3--added Bi0.5Na0.5TiO3 materials was enhanced to around 39.6 memu/g (around 18.7

memu/g after substracting the paramagnetic-like component) at 6 kOe. This value is higher than

that of Bi0.5Na0.5TiO3 materials doped with single Mn (~9 memu/g), Cr (~1.5 memu/g), Fe (~11

memu/g), and Co (~3 memu/g) [69]. Coey et al. pointed out that 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.

6
IV. Conclusion

MgFeO3- materials existed as a well solid solution in Bi0.5Na0.5TiO3 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 ferroelectric materials to fabricate green electronic

devices.

V. Acknowledgment

This work was financially supported by The Ministry of Education and Training, Viet Nam,
under project number B2016-BKA-25.

7
Figure Captions

Fig. 1. (a) X-ray diffraction pattern of MgFeO3- solid solution into Bi0.5Na0.5TiO3 with various
concentrations within the 2 range of 20 to 70; (b) Magnified X-ray diffraction within 2
range of 3134 for comparing setline (012)/(110) peaks; (c) Raman scattering spectra of
MgFeO3- solid solution into Bi0.5Na0.5TiO3 with various concentrations at 200 cm1 to 1000
cm1; and (d) deconvolution Raman peaks of pure Bi0.5Na0.5TiO3, Bi0.5Na0.5TiO3-3MgFeO3-.
and Bi0.5Na0.5TiO3-5MgFeO3-.

Fig. 2. (a) UV-Vis absorption spectra of MgFeO3--modified Bi0.5Na0.5TiO3 samples as a


function of MgFeO3- concentration; and (b) the (h)2 proposal with photon energy (h) of
Bi0.5Na0.5TiO3 samples as a function of the amount of MgFeO3- 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- added.

Fig. 3. MH curves of the pure Bi0.5Na0.5TiO3 and MgFeO3--modified Bi0.5Na0.5TiO3 samples


added with various amounts of MgFeO3- at room temperature.

8
References

[1] N. D. Quan, L. H. Bac, D. V. Thiet, V. N. Hung, and D. D. Dung, Adv. Mater. Sci. Eng.
2014 (2014) article ID 365391.

[2] P. Baettig, C. F. Schelle, R. Lesar, U. V. Waghmare, and N. A. Spaldin, Chem. Mater.


17 (2005) 13761380.

[3] X. He, and K. J. Jin, Phys. Rev. B 94 (2016) 224107.

[4] G. A. Smolensky, V. A. Isupov, A. I. Agranovskaya, and N. N. Krainic, Fizika Tverdogo


Tela. 2 (1960) 29822985.

[5] T. Takenaka, K. I. Maruyama, and K. Sakata, Japanese J. Appl. Phys. 30 (1991) 2236-2239.

[6] Y. Wang, G. Xu, L. Yang, Z. Ren, X. Wei, W. Weng, P. Du, G. Shen, G. Shen, and G. Han,
Mater. Sci. Poland 27 (2009) 471-476.

[7] Y. Wang, G. Xu, X. Ji, Z. Ren, W. Weng, P. Du, G. Shen, and G. Han, J. Alloys Compound.
475 (2009) L25-L30.

[8] L. T. H. Thanh, N. B. Doan, L. H. Bac, D. V. Thiet, S. Cho, P. Q. Bao, and D. D. Dung,


Mater. Lett. 186 (2017) 239-242.

[9] L. T. H. Thanh, N. B. Doan, N. Q. Dung, L.V. Cuong, L. H. Bac, N. A. Duc, P. Q. Bao, and
D. D. Dung, J. Electron. Mater. 46 (2017) 3367-3372.

[10] J. M. D. Coey, M. Venkatesan, and C. B. Fitzgerals, Nature Mater. 4 (2005) 173-179.

[11] F. Guo, B. Yang, S. Zhang, F. Wu, D. Liu, P. Hu, Y. Sun, D. Wang, and W. Cao, Appl.
Phys. Lett. 103 (2013) 182906.

9
[12] A. Ullah, C. W. Ahn, K. B. Jang, A. Hussain, and I. W. Kim, Ferroelectrics 404 (2010)
167-172.

[13] J. S. Kim, C. H. Chung, H. S. Lee, and S. T. Chung, J. Korean Phys. Soc. 58 (2011) 659-
662.

[14] W. Hume-Rothery, Atomic Theory for Students of Metallurgy, The Institute of Metals,
London, 1969 (fifth reprint).

[15] C. Barry Carter, and M. Grant Norton, Ceramic Materials: Science and Engineering,
Springer, 2007, pp.126.

[16] R. D. Shannon, Acta. Cryst. A 32 (1976) 751-767.

[17] M. K. Niranjan, T. Karthik, S. Asthana, and J. Pan, J. Appl. Phys. 113 (2013) 194106.

[18] J. Kreisel, A. M. Glazer, P. Bouvier, and G. Lucazeau, Phys. Rev. B 63 (2001) 174106.

[19] Y. Chen, K. H. Lam, D. Zhou, J. Y. Dai, H. S. Luo, X. P. Jiang, and H. L. W. Chan, Inter.

Ferroelectric 141 (2013) 120-127.

[20] J. Suchanicz, I. J. Sumara, and T. V. Kruzina, J. Electroceram. 27 (2011) 45-50.

[21] D. D. Dung, D. V. Thiet, D. Odkhuu, L. V. Cuong, N. H. Tuan, and S. Cho, Mater. Lett.

156 (2015) 129-133.

[22] N.D. Quan, V.N. Hung, N.V. Quyet, H.V. Chung, and D.D. Dung, AIP Advances 4 (2014)

017122.

[23] N.V. Quyet, L.H. Bac, D. Odkhuu, and D.D. Dung, J. Phys. Chem. Solids 85 (2015) 148-

154.

[24] L. H. Bac, L. T. H. Thanh, N. V. Chinh, N. T. Khoa, D. V. Thiet, T. V. Trung, and D. D.

Dung, Mater. Lett. 164 (2016) 631-635.

10
[25] Y. Zhang, J. Hu, F. Gao, H. Liu, and H. Qin, Comput. Theor. Chem. 967 (2011) 284-288.

[26] K. Balamurugan, N. H. Kumar, J. A, Chelvane, and P.N. Santhosh, Physica B 407 (2012)

2519-2523.

11
Hung et al. Fig. 1

12
Hung et al. Fig. 2

13
Hung et al. Fig. 3

14
Highlights

- The Bi0.5Na0.5TiO3-MgFeO3- materials were fabricated by solgel method.

- The MgFeO3- materials existed as a well solid solution and possessed the structure of

Bi0.5Na0.5TiO3.

- The room-temperature ferromagnetism in Bi0.5Na0.5TiO3-MgFeO3- samples was

obtained.

- Our work was provided new method to inject room temperature ferromagnetism in lead-

free ferroelectric materials.

15

Anda mungkin juga menyukai