Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Vacuum 191 (2021) 110389

Contents lists available at ScienceDirect

Vacuum
journal homepage: www.elsevier.com/locate/vacuum

Short communication

Tunable optical properties of Bi1/2Na1/2TiO3 materials via Sm1/2Na1/


2TiO3 addition

Nguyen Hoang Thoan a, Nguyen Huu Lam a, Ha Thi Thu Hieu b, c, Nguyen The Hoang a,
Pham Van Vinh d, Nguyen Van Hao b, Nguyen Ngoc Trung a, Duong Quoc Van d, **,
Dang Duc Dung a, *
a
Multifunctional Ferroics Materials Lab., School of Engineering Physics, Ha Noi University of Science and Technology, 1 Dai Co Viet road, Ha Noi, Viet Nam
b
Institute of Science and Technology, TNU - University of Sciences, Tan Thinh ward, Thai Nguyen City, Viet Nam
c
Bac Kan Center for Regular and Vocational Education, Bac Kan, Viet Nam
d
Faculty of Physics, Ha Noi National University of Education , 136 Xuan Thuy Street, Cau Giay District, Hanoi, Viet Nam

A R T I C L E I N F O A B S T R A C T

Keywords: In this work, the Sm1/2Na1/2TiO3-addition Bi1/2Na1/2TiO3 materials as solid solutions were synthesized via a
Bi1/2Na1/2TiO3 chemical route method. The X-ray diffraction and Raman scattering results indicated that Sm1/2Na1/2TiO3 ma­
Sm1/2Na1/2TiO3 terials were well solute into the host Bi1/2Na1/2TiO3 materials to form solid solutions. The optical bandgap of the
Sol-gel
pure Bi1/2Na1/2TiO3 materials was estimated from ultraviolet–visible spectroscopy of 3.07 eV. The bandgap
Photoluminescence
Solid solution
slightly reduced to 2.94 eV as increasing the Sm1/2Na1/2TiO3 amounts up to 9 mol.%. Strong photoluminescence
(PL) of Sm1/2Na1/2TiO3-addition Bi1/2Na1/2TiO3 materials was obtained in the visible wavelength range of
550–750 nm which related to photoluminescence of Sm cations randomly incorporated into the host lattice of
Bi1/2Na1/2TiO3 materials. Our work was promised to show a new method to integrate the PL properties in lead-
free ferroelectric materials.

1. Introduction field-induced strain Smax/Emax of 88 p.m./V [7,8]. The Bi1/2Na1/2TiO3


materials exhibited weak-ferromagnetic properties at room temperature
Lead-free ferroelectric Bi1/2Na1/2TiO3 materials had been attached due to the presence of various types of defects which were formed during
much attention because of their ability to replace Pb(Zr,Ti)O3-based sample fabrication [4,9]. The Bi1/2Na1/2TiO3 materials exhibited direct
materials, which contain a large amount of Pb toxic elements to human transition with an optical bandgap of approximately 3.02–3.11 eV,
health and pollution protection [1]. In addition, ongoing research had depending on fabricating conditions and chemical defects [4,10].
tried to extend the function of lead-free ferroelectric materials for Thanks to the well-solid solution, the properties of Bi1/2Na1/2TiO3
electro-optical applications that used ferroelectric and piezoelectric materials were greatly enhanced by using various perovskite materials
properties of their materials [2,3]. Recently, the properties of lead-free [7,12–17]. By co-addition of BaTiO3 and Bi2FeCrO6, the EC value of
ferroelectric Bi1/2Na1/2TiO3 materials had been approved to extend Bi1/2Na1/2TiO3 materials was reduced to 28.7 kV/cm while Pr and d33
functional materials for smart material application [4]. The ferroelectric values were increased to 42.3 μC/cm2 and 214 pC/N, respectively [12].
Bi1/2Na1/2TiO3 materials, first discovered by Smolenski et al., exhibited The Pr and spontaneous polarization (PS) of the Bi1/2Na1/2TiO3 materials
strong ferroelectric properties such as remanent polarization (Pr) ~ 38 increased to 48.5 μC/cm2 by Ba(Mg1/3Nb2/3)O3-modification [13]. The
μC/cm2, coercive field (EC) of 7.30 kV/mm, and a high Curie tempera­ dielectric constant and piezoelectric constant d33 of the Bi1/2Na1/2TiO3
ture (TC) of ~320 ◦ C [5,6]. However, the high EC values prevented the materials increased to 144 pC/N and 893 pC/N, respectively, by adding
transfer of the Bi1/2Na1/2TiO3 materials to actual industrial application Bi1/2K1/2TiO3 [14]. Rahman et al. reported that the electrical
because a high electrical field required for polarizing process resulted in field-induced strain of the Bi1/2Na1/2TiO3 materials increased up to 500
poor piezoelectric properties of d33 ~ 58 pC/N and low electrical p.m./V by 5.5 mol.%-BaZrO3 modification [8]. Shin et al. reported that

* Corresponding author.
** Corresponding author.
E-mail addresses: vandq@hnue.edu.vn (D.Q. Van), dung.dangduc@hust.edu.vn (D.D. Dung).

https://doi.org/10.1016/j.vacuum.2021.110389
Received 6 May 2021; Received in revised form 31 May 2021; Accepted 11 June 2021
Available online 15 June 2021
0042-207X/© 2021 Elsevier Ltd. All rights reserved.
N.H. Thoan et al. Vacuum 191 (2021) 110389

the (Sr0.7Bi0.2)TiO3-modified Bi1/2Na1/2TiO3-based materials exhibited magnetic stirring until turning transparent. Then, NaNO3 and Sm
strong enhancement of electrostrictive strain up to 0.152% and elec­ (NO3)3.6H2O were added. The solutions were continuously kept under
trostrictive coefficient of 0.0297 m4/C2 [15]. In addition, Liu et al. re­ magnetic stirring for about 1 h. Thus, the acetylacetone was dropped to
ported that properties of Bi1/2Na1/2TiO3-based materials slightly doped the solutions with the volume ratio of acid acetic and acetylacetone of
with cerium and stannum could be improved by optimizing the sintering 1:1. Subsequently, the C12H28O4Ti were taken to drop into the solution.
processes [16,17]. In addition, the magnetic properties of Bi1/2Na1/2 The solutions were continuously kept under magnetic stirring for about
TiO3 materials were strongly enhanced by doping various ABO3-type 3–5 h to obtain homogeneous sols. The sols were dried in the oven at
materials such as CaFeO3-δ, BaFeO3-δ, SrMnO3-δ, BaCoO3-δ, NiTiO3, and 100 ◦ C to form dry gels. The gels were simply filtered and annealed at
Bi(Ti1/2Fe1/2)O3 [4,18–22]. In other words, the properties of 900 ◦ C for 5 h to form powder samples. To prevent the loss of Na during
Bi1/2Na1/2TiO3 materials were greatly enhanced via using various types gelling and sintering processes, NaNO3 was weighed to an extra amount
of impurity materials. of about 30–50 mol.% [18–22,27–29]. The chemical composition of
Recently, there was a new trend in advanced functional materials via pure Bi1/2Na1/2TiO3 and Sm1/2Na1/2TiO3-modified Bi1/2Na1/2TiO3 ma­
a combination of photoluminescence (PL) behavior of rare-earth-doped terials was qualified by energy-dispersive X-ray spectroscopy (EDS)
lead-free ferroelectric materials [23–25]. Therefore, enhancement of the measurements. The crystal structural symmetry of pure Bi1/2Na1/2TiO3
PL properties of Bi1/2Na1/2TiO3 materials promised to create a new type and Sm1/2Na1/2TiO3-modified Bi1/2Na1/2TiO3 materials was determined
of electronic device. The Bi1/2Na1/2TiO3 materials are ferroelectric ones by X-ray diffraction (XRD) spectroscopy. The vibration modes of sam­
[1,4–8]. However, inside ferroelectric materials, separated electron-hole ples were measured by using Raman scattering. The optical properties of
pairs were hard to recombine, resulting in low-efficiency photo­ pure Bi1/2Na1/2TiO3 material and Sm1/2Na1/2TiO3 -modified
luminescence [26]. Therefore, the photoluminescence of the Bi1/2Na1/2TiO3 materials were measured by Ultraviolet–visible spec­
Bi1/2Na1/2TiO3 materials was most related to surface defects because troscopy (UV–Vis) and photoluminescence spectroscopy (PL)
there were fewer bonding pairs at the surface than inside of materials [4, measurement.
18,19,26]. Doping of transition metals and/or solid solutions with
dopants containing transition metals resulted in suppression of the PL 3. Results and discussions
properties of the Bi1/2Na1/2TiO3 materials [27–29]. Recently, the elec­
trical properties of lead-free Bi1/2Na1/2TiO3-based materials were re­ The crystal structure of pure Bi1/2Na1/2TiO3 and Sm1/2Na1/2TiO3-
ported to be strongly enhanced via the addition of rare-earth materials modified Bi1/2Na1/2TiO3 materials with various Sm1/2Na1/2TiO3 con­
[17,30,31]. The PL performance of lead-free ferroelectric Bi1/2Na1/2 centrations was characterized by the X-ray diffraction patterns, as
TiO3 materials could be improved via using rare-earth as impurities, e.g. shown in Fig. 1 (a). The XRD results showed that pure Bi1/2Na1/2TiO3
Sm, Ho, Er, Eu, Nd, Pr, etc. [32–37]. The rare-earth doping in material exhibited a rhombohedral structure because all peak positions
Bi1/2Na1/2TiO3 materials resulted in a strong enhancement of the and their relative intensities were well matched with JCPDF card. With
luminescence emission which was related to PL emission from 4f-to-4f the addition of Sm1/2Na1/2TiO3 materials, the materials exhibited follow
levels of rare-earth cations. the structure of the host Bi1/2Na1/2TiO3 materials. No trade phase and/
Sm1/2Na1/2TiO3 materials were first fabricated by a solid-state re­ or phase segregation could be observed under X-ray diffraction patterns,
action method by Barik et al. [38]. The Sm1/2Na1/2TiO3 materials had an revealing that the Sm1/2Na1/2TiO3 materials were well solute into host
orthorhombic crystal structure with the parameters a = 3.8263(7) Å, b Bi1/2Na1/2TiO3 materials. The role of Sm1/2Na1/2TiO3 solute phase in
= 3.8124(7) Å and c = 3.8425(7) Å where Sm and Na cations randomly Bi1/2Na1/2TiO3 materials was shown in Fig. 1 (b), where the X-ray
distributed at A-sites in the perovskite structure [38]. The dielectric diffraction patterns were magnified in the 2θ-range of 31.0–34.0◦ . The
properties of the Sm1/2Na1/2TiO3 materials were reported to be satellite (012)/(110) peaks were overlapped together which were orig­
enhanced by filling polytetrafluoroethylene or doping with Cr2O3, Sm inated from the symmetry of rhombohedral structure [18–22]. The peak
(Mg0.5Ti0.5)O3, or (Al, Ta) co-doping [39–41]. Recently, the magnetic positions of Bi1/2Na1/2TiO3 materials trended to shift to higher 2θ angles
properties of Bi1/2Na1/2TiO3 materials were reported to be strongly as increasing the Sm1/2Na1/2TiO3 amounts. The results provided solid
enhanced via modification of impurity phases containing transition evidence for lattice parameter distortion of the host Bi1/2Na1/2TiO3
metals as solid solutions [4,18–22,27–29]. Herein, materials containing materials during the solid solution process of the Sm1/2Na1/2TiO3 phase.
Sm cations as 4f materials were expected to luminesce. However, up to The XRD peak positions of the pure Bi1/2Na1/2TiO3 and Sm1/2Na1/2
date, there was no report on the optical properties of Sm1/2Na1/2TiO3 TiO3-modified Bi1/2Na1/2TiO3 materials were distinguished by the Lor­
materials modified Bi1/2Na1/2TiO3 materials. entzian fitting with r-square higher than 0.99, as shown in red and blue
In this work, a new solid solution (1-x)Bi1/2Na1/2TiO3 +xSm1/2Na1/ lines for (012) and (110) peaks, respectively. The fit results were
2TiO3 system was well synthesized via a simple chemical method. The interesting and showed that the intensities and positions of the double
structure of Sm1/2Na1/2TiO3 modified Bi1/2Na1/2TiO3 materials was a (012)/(110) peaks of the host Bi1/2Na1/2TiO3 materials were very
distortion of rhombohedral symmetry. The optical bandgap of Sm1/ complex distorted, depending on the amounts of Sm1/2Na1/2TiO3 solute
2Na1/2TiO3-modified Bi1/2Na1/2TiO3 materials was estimated in the in the host Bi1/2Na1/2TiO3 materials. Furthermore, the lattice parame­
range of 3.07–2.94 eV depending on Sm1/2Na1/2TiO3 concentrations in ters a and c of the pure Bi1/2Na1/2TiO3 and Sm1/2Na1/2TiO3-modified
the solid solutions. Strong photoemission in the wavelength range of Bi1/2Na1/2TiO3 materials were estimated and shown in Fig. 1(c) as a
550–750 nm was obtained from the Bi1/2Na1/2TiO3 materials as a function of Sm1/2Na1/2TiO3 concentration. The results showed that
modification by Sm1/2Na1/2TiO3 addition. distorted lattice parameters of Bi1/2Na1/2TiO3 compounds were not
linear as a function of Sm1/2Na1/2TiO3 concentrations, which has com­
2. Experimental plexed distortion in lattice parameters. The complex distortion of the
lattice parameters of the host Bi1/2Na1/2TiO3 materials via
In this study, (1-x)Bi1/2Na1/2TiO3 +xSm1/2Na1/2TiO3 materials with Sm1/2Na1/2TiO3 addition could be explained by the radius difference of
x = 0, 0.5, 1, 3, 5, 7, and 9 mol.%; named as pure BNT and BNT-xSm, the impurity cations in the host matrix. Based on Shannon’s report, the
were synthesized by a simple chemical method. The starting materials radius of Bi3+ cations (in the coordination of VIII) and Na+ cations (in
included bismuth nitrate pentahydrate (Bi(NO3)3⋅5H2O), sodium nitrate the coordination of XII) were 1.17 Å and 1.39 Å, respectively [42]. The
(NaNO3), samarium nitrate hexahydrate (Sm(NO3)3.6H2O), and tetrai­ radius of Sm3+ cations (in the coordination of XII) was 1.24 Å [42].
sopropoxytitanium (IV, C12H28O4Ti). First, Bi(NO3)3⋅5H2O was weighed Based on the Hume–Rothery rules, Sm cations enter at the substituted
and added in the solutions of acetic acid and de-ions water with the A-site of Bi1/2Na1/2TiO3 materials because of the − 3.13% difference
volume ratio of VH2O:VCH3COOH = 5:1. The solutions were kept under between the radius of Sm cations and the average radius of

2
­
­
­
N.H. Thoan et al. Vacuum 191 (2021) 110389

Fig. 1. (a) XRD patterns of pure Bi1/2Na1/2TiO3 and Sm1/2Na1/2TiO3-modified Bi1/2Na1/2TiO3 compounds with various of Sm1/2Na1/2TiO3 concentrations, (b) the
magnification with deconvolution peaks of XRD pattern in 2θ range from 31.0◦ to 34.0◦ , and (c) the (a,c) lattice parameters of Bi1/2Na1/2TiO3 as a function of the
Sm1/2Na1/2TiO3 concentration.

(Bi1/2Na1/2)2+ [43–45]. The difference between the impurity Sm cations Bi1/2Na1/2TiO3 crystal was possibly reduced because of oxygen va­
and the host Bi and Na cations were +5.98% and − 10.79%, respectively. cancies nearby, where the size 1.31 Å of oxygen vacancies was smaller
Therefore, the random incorporation of Sm3+ cations into A-sites in the than that of anion oxygen of 1.4 Å [51]. The flaccidity of the size of O
Bi1/2Na1/2TiO3 perovskite structures resulted in a complex distortion of vacancies on the lattice parameters has a more significant influence than
lattice parameters. The substitution of Sm3+ cations in Bi3+-sites resul­ that of dopants in perovskite Bi1/2Na1/2TiO3-based materials [4,18–22].
ted in compression of lattice parameters because the radius of Sm3+ Note that the oxygen vacancies were created via the unbalance valence
cations is smaller than that of Bi3+ cations. However, the incorporation state between Sm3+ and Ti4+ cations. Herein, we needed to note that the
of Sm3+ cation into Na + -sites resulted in an expansion of lattice pa­ oxygen bounding surrounds the Ti4+ cations were possibly reduced the
rameters because of the larger radius of Sm3+ cations in comparison valence state of Ti cations from the Ti4+ state to the Ti3+ state [4,18,52].
with Na + ones at A-sites. In addition, the difference between the valence We recently reported that Ti3+-defects were stable in Bi1/2Na1/2TiO3-­
states of Sm3+ impurity cations and Na + cations during incorporation based materials [18]. The reduction of the valence state of Ti4+ to Ti3+
with the host Bi1/2Na1/2TiO3 lattices might create Na-vacancies. The resulted in local chemical tensile strain because the size 0.670 Å of Ti3+
stable Na-vacancies in the host crystal Bi1/2Na1/2TiO3 structure resulted cations was larger than 0.605 Å of Ti4+ cations [42]. At this moment, we
in a distortion of lattice parameters, resulting in a reduction of the lattice were unable to provide the main reason for the observation of lattice
strain energy of host lattice Bi1/2Na1/2TiO3 materials [46–48]. In addi­ parameter distortion of the host Bi1/2Na1/2TiO3 materials via
tion, Luo et al. reported that the Er cations were possibly doped at both Sm1/2Na1/2TiO3 addition as numerous parameters contributed to the
A-sites (Bi-, Na-sites) and B-sites (Ti-sites) in perovskite structure of suggestion. However, the observation of compression of lattice param­
Bi1/2Na1/2TiO3 materials [49]. A similar effect was obtained in Eu doped eters of Bi1/2Na1/2TiO3 materials via a solid solution of Sm1/2Na1/2TiO3
Bi1/2Na1/2TiO3 materials [50]. Taking account of our results, if we as impurities provided strong evidence of random incorporation of im­
considered the Sm3+ cations randomly filled to Ti-site in Bi1/2Na1/2TiO3 purity cations into the host Bi1/2Na1/2TiO3 lattices.
crystal during the solid solution, the distortion lattice parameter of host Fig. 2 (a) shows the Raman spectra of the pure Bi1/2Na1/2TiO3 and
Bi1/2Na1/2TiO3 materials became more complex. According to Shan­ Sm1/2Na1/2TiO3-modified Bi1/2Na1/2TiO3 materials with various Sm1/
non’s report, the size of Sm3+ cations at coordination of IV was 0.958 Å 2Na1/2TiO3 amounts at room temperature. The Raman spectrum of the
which was larger than that of Ti4+ cations of 0.605 Å with the same pure Bi1/2Na1/2TiO3 exhibited three overlapping humps. Unclear peaks
coordination, resulting in expansion of the lattice parameters of the host in the Raman scattering spectrum of pure Bi1/2Na1/2TiO3 materials came
Bi1/2Na1/2TiO3 crystals [42]. The differences between the radius of from a random distribution of Na and Bi cations at A-sites in the
Sm3+ cations and those of Ti4+ cations substituted at the Ti-site of rhombohedral symmetry [53,54]. Our observation on Raman scattering
Bi0.5Na0.5TiO3 materials are 58.35%, which, according to the results of the pure Bi1/2Na1/2TiO3 was well consistent with recent arti­
Hume–Rothery rules, is too large to allow replacement because of the cles for phonon scattering modes of Bi1/2Na1/2TiO3 materials synthe­
increased lattice energy [43–45]. The large difference between the sizes sized by sol-gel technique [18–22]. The Raman spectra of the
of the impurity Sm3+ cations and the host Ti4+ cations induced a large Sm1/2Na1/2TiO3-addition Bi1/2Na1/2TiO3 materials were similar in
local tensile strain in the crystal. However, the local strain in shape with the pure BNT, further confirming the same crystal structure

3
N.H. Thoan et al. Vacuum 191 (2021) 110389

Fig. 2. (a) Raman scattering spectra and (b) the deconvolution of Raman scattering peaks of pure Bi1/2Na1/2TiO3 and Sm1/2Na1/2TiO3-modified Bi1/2Na1/2TiO3
compounds with various Sm1/2Na1/2TiO3 amounts as solid solutions.

the addition of Sm1/2Na1/2TiO3. Furthermore, the Raman spectra of cm− 1 and 300 cm− 1 were associated with the Ti–O vibrations, while the
Sm1/2Na1/2TiO3-modified Bi1/2Na1/2TiO3 materials did not show any high-frequency bands at about 490 cm− 1, 530 cm− 1, and 610 cm− 1 were
extra peaks, indicating that no extra phase or impurity-related phase dominated by the TiO6-octahedra vibration and oxygen displacement
existed in our studied samples. The results were well consistent with the [56]. He et al. pointed out that the overlapping Raman spectra of
single-phase crystal structure which had been confirmed by the XRD Bi1/2Na1/2TiO3-based materials could be divided into four main regions,
study. The Raman peaks of pure Bi1/2Na1/2TiO3 and Sm1/2Na1/2TiO3-­ including: i) the modes in the wavenumber range below 200 cm− 1 are
modified Bi1/2Na1/2TiO3 materials were distinguished by Lorentzian assigned to the A-site vibration of perovskite oxides, ii) the modes in the
fitting with r-square higher than 0.99, as shown in Fig. 2 (b). Nice peaks wavenumber range of 200–450 cm− 1 are associated with the Ti–O vi­
were obtained in Raman spectra in the frequency range of 300–950 brations, iii) the modes in the range of 450–700 cm− 1 are related to the
cm− 1. A smaller number of modes than theoretical predictions was vibrations of TiO6 octahedra, and iv) the modes with the wavenumber
observed, which was possibly related to phonon-frequency degradation higher than 700 cm− 1 are associated to longitudinal optical vibrations of
or insufficient intensities arising from the small polarizability of several A1 and E overlapping bands [57]. Our fitting results pointed that the
modes [53,54]. The Raman peak positions were well consistent with Raman peaks at approximately 402 and 585 cm− 1 shifted to higher
theoretical prediction and recent experimental results [18–22,53,54]. frequencies as increasing the Sm1/2Na1/2TiO3 amounts in the host
Lui et al. reported that the mode at 270 cm− 1 of Bi1/2Na1/2TiO3-based Bi1/2Na1/2TiO3 materials. The shift of Raman scattering peaks was
materials was dominated by A1 mode assigned to Ti–O vibrations, while possibly related to distortion of the Bi1/2Na1/2TiO3 structure due to
the modes in the mid-wavenumber region of 450–700 cm− 1 related to random distribution of impurity cations during Sm1/2Na1/2TiO3
the vibration of TiO6 octahedra, and the modes in the high-wavenumber solid-solution process into the host Bi1/2Na1/2TiO3 crystal.
region above 700 cm− 1 were linked to the overlapping bands A1(LO) Fig. 3 (a) showed the UV–Vis spectra of pure Bi1/2Na1/2TiO3 and
and E(LO) [55]. Liu et al. also reported that the modes near 100 cm− 1 of Sm1/2Na1/2TiO3-modified Bi1/2Na1/2TiO3 materials. The spectrum of
Bi1/2Na1/2TiO3-based materials were related to vibration involving the the pure Bi1/2Na1/2TiO3 material exhibited a single edge transition and
perovskite A-sites with A1 symmetry, and the broad bands near 240 was tailored with a slight tail. The absorption edge of the pure Bi1/2Na1/

Fig. 3. (a) UV–Vis spectra, and (b) the plots of (αhν)2 as functions of photon energy (hν) of pure Bi1/2Na1/2TiO3 compounds and Sm1/2Na1/2TiO3-modified Bi1/2Na1/
2TiO3 compounds. The inset showed the dependence of Eg values of Bi1/2Na1/2TiO3 as a function of Sm1/2Na1/2TiO3 amounts.

4
N.H. Thoan et al. Vacuum 191 (2021) 110389

2TiO3 material was estimated at approximately 418 nm, the tail was Fig. 4 (a) showed the PL emission spectra of Sm1/2Na1/2TiO3-modi­
prolonged to 580 nm. These results were well consistent with recently fied Bi1/2Na1/2TiO3 materials with different Sm1/2Na1/2TiO3 concen­
obtained absorption spectra of pure Bi1/2Na1/2TiO3 materials fabricated trations under an excitation wavelength of 475 nm. The results were
by the sol-gel method [4,10,18–22,58]. The appearance of a slight tail in important to point out that the strong emission was obtained in wide
the absorption spectrum of the Bi1/2Na1/2TiO3 materials was originated wavelength regions. The PL of Sm1/2Na1/2TiO3-modified Bi1/2Na1/2TiO3
from self-defects such as vacancies of Bi, Na, Ti, or O, and/or materials overcame the PL of the pure Bi1/2Na1/2TiO3 materials [4,18,
Ti3+-defects [4,22]. The addition of Sm1/2Na1/2TiO3 to the host 22,26]. The Sm1/2Na1/2TiO3 addition into host the Bi1/2Na1/2TiO3 ma­
Bi1/2Na1/2TiO3 materials caused a slight shift in the edge transition, terials exhibited strong photoluminescence which overcame the sup­
indicating that the optical bandgap (Eg) of the host Bi1/2Na1/2TiO3 pression of photoluminescence intensity of the host Bi1/2Na1/2TiO3
materials was tuned as a function of Sm1/2Na1/2TiO3 concentrations in materials via the addition of transition metals [4,18,22]. The results
the solid solutions. In addition, it was suggested that the appearance of a were also important to indicate that the random incorporation of
hump around 450–510 nm, as shown in the inset of Fig. 3 (a), was rare-earth Sm3+ cations into host Bi1/2Na1/2TiO3 materials exhibited
related to the transition of electrons from 6H5/2 to 4I11/2 and 4M15/2 stronger photoluminescence than that of self-defect and/or
levels [59]. In recent reports, the UV–Vis absorption edge of surface-defect-mediated photoluminescence of lead-free ferroelectric
Bi1/2Na1/2TiO3 materials tended to shift to a higher absorption wave­ materials [4,18,22,26]. PL emission of rare-earth Sm3+ cations doped in
length as increasing the dopant amounts, e.g. Fe, Co, Ni, Mn, and Cr [11, the lead-free ferroelectric Bi1/2Na1/2TiO3 materials in this study was
60–63]. However, unlike single transition-metal dopants, the influence stronger than that of the same materials doped with transition metals [4,
of Sm cations on the absorption edge was complex, which was strongly 18,22]. As shown in Fig. 4 (a), four main peaks in the visible region
dependent on the Sm amounts randomly incorporated with the host corresponded to intra-4f transitions of Sm3+ cations from 4G5/2 level to
6
Bi1/2Na1/2TiO3 lattices. Recently, we predicted that the electronic band H5/2, 6H7/2, 6H9/2, and 6H11/2 levels. These results agreed with the ones
structural of Bi1/2Na1/2TiO3 materials consisted of Bi-6p and Ti-3d or­ recently reported for Sm3+ emission transitions in the visible region [32,
bitals for building the valence band while the conduction band was 71–74]. Furthermore, the PL intensity as a function of Sm1/2Na1/2TiO3
composed of O-2p orbitals where the direction transition was suggested concentrations was compared with strong photoemission peaks. Fig. 4
to charge transfer from the critical point Γ(0, 0, 0) in the reciprocal space (b) showed the relative PL intensity of the strongest peak (at approxi­
[64]. Zeng et al. reported that the optical properties of Bi1/2Na1/2TiO3 mately 601 nm, 4G5/2 → 6H7/2) as a function of Sm3+ concentrations in
materials were possibly determined by charge-transfer transitions from our studied samples. The 4G5/2 → 6H7/2 emission intensity increased
O-2p to Ti-3d or Bi-6p states [65]. Thus, the Wood-Tauc method was with increasing Sm3+ concentrations in the lower content region until
used to estimate Eg values, where the Eg values of the Bi1/2Na1/2TiO3 getting saturation for the sample with 5 mol.% Sm3+, then decreased
materials were obtained from extrapolation to zero of the linear fittings slightly. The decrease could be explained by concentration-quenching
of (αhγ)2 vs. photon energy (hγ) plots, as shown in Fig. 3 (b) [4,10,22]. processes: a Sm3+ doping level higher than 5 mol.% caused a decrease
The pure Bi1/2Na1/2TiO3 materials exhibited the Eg value of about 3.07 of the distance between Sm3+ cations and non-radiative energy transfer
eV, which was well consistent with recently observed Eg values of from one PL activator to another activator, leading to the lowering of the
Bi1/2Na1/2TiO3 materials synthesized by sol-gel technique [4,10,11, PL intensity [32]. Interestingly, most reports on the PL influence of Sm3+
60–63]. The addition of Sm1/2Na1/2TiO3 to the host Bi1/2Na1/2TiO3 doped lead-free ferroelectric Bi1/2Na1/2TiO3-based materials, such as
materials resulted in a reduction in the Eg values. However, Eg values of pure Bi1/2Na1/2TiO3, SrTiO3–Bi1/2Na1/2TiO3, BaTiO3–Bi1/2Na1/2TiO3
Bi1/2Na1/2TiO3 materials did not simply follow in one direction as materials, focused on the up-conversion; but the relation between Sm3+
increasing the Sm1/2Na1/2TiO3 concentrations. In fact, the Eg values of emission and excitation through the host was not well reported [32,75,
Bi1/2Na1/2TiO3 materials were very complex distorted and slightly 76]. Recently, we provided that the Bi1/2Na1/2TiO3 materials were
reduced to 2.94 eV for 9 mol.% Sm1/2Na1/2TiO3 solid solution. Our re­ exhibited photoluminescence in the range from 476 to 510 nm where
sults were well consistent with recent observations of the influence of maximal emission was obtained around 489 nm [4,18,22]. The
rare-earth-doped lead-free ferroelectric Bi1/2Na1/2TiO3 materials such multi-emission peaks observed in the wavelength from 476 to 510 nm
as Ce, Nd, Pr, etc. [66–68]. The optical bandgap values of the were most related to self-defect and/or surface defects rather than
Bi1/2Na1/2TiO3 materials as a function of the Sm1/2Na1/2TiO3 amounts electron transitions from the conduction band to the valence band, as
materials were affected by many parameters. First, the random incor­ considering optical bandgap values [4,18,22]. Note that the
poration of Sm3+ cations at A-sites and B-sites in the host Bi1/2Na1/2TiO3 Bi1/2Na1/2TiO3 materials are ferroelectric materials, which existed nat­
materials resulted in creating various new localized states of Sm cations ural domain polarization, preventing the photo-hole and
which were dependent on the substituted site [69]. Secondary, Na- and photon-electron recombination to generate photoemission [26]. There­
O-vacancies, which were created due to random multi-site substitution fore, the emission peaks of lead-free ferroelectric materials were mostly
of Sm3+ cations for Na + -sites or Ti4+-sites, also resulted in a formation related to defects rather than photoemission peaks from band-to-band
of new local states of self-defects in the electronic band structure of the transitions [4,18,22,26]. Recently, the photoemission of Bi1/2Na1/2
host Bi1/2Na1/2TiO3 materials [4,22]. Moreover, the possible valence TiO3 materials was suppressed by doping the transition metals [4,18,
state of Ti4+ cations might reduce to the Ti3+ state via oxygen vacancies 22]. The observed results pointed out that the local defects of impurity
bounding around, resulting in new local states in the electronic band transition metals in the electronic band structure absorbed photons
structure of the Bi1/2Na1/2TiO3 materials [4,22]. In addition, Behara emitted from natural defects of the host Bi1/2Na1/2TiO3 materials,
et al. predicted that compression strain led to an expansion of the optical resulting in suppression of photoemission intensity. In other words, the
bandgap of Bi1/2Na1/2TiO3 materials while tensile strain resulted in local defects of transition metal impurities made the trapping of
lower bandgap values [70]. In other words, many parameters affected photoemission of natural defects of the host Bi1/2Na1/2TiO3 materials.
the electronic band structure of Bi1/2Na1/2TiO3 via Sm1/2Na1/2TiO3 Recently, Rosales et al. proposed that the possible secondary transitions
addition as solid solutions. In fact, at this moment, the complex of photons from the conduction band to an inter-band level of Yb3+
dependence of optical band gap energy of Bi1/2Na1/2TiO3 materials on cations resulted in photoluminescence activation [77]. Therefore, we
Sm1/2Na1/2TiO3 concentrations was not well understood and need suggested that the possible transitions from natural defects to a Sm3+
further theoretical investigation. However, the observation of tunable inter-band level were active for photoluminescence of 4f-4f level tran­
the optical band gap energy of the host Bi1/2Na1/2TiO3 materials via sitions. However, the detail of nature defects in Sm3+ cations in host
Sm1/2Na1/2TiO3 addition provided important indirect evidence for Bi1/2Na1/2TiO3 materials needed further study via theoretical
random incorporation of impurity cations during the formation of the calculation.
solid solutions. Fig. 5 showed the magnifications of the PL spectra of the pure Bi1/

5
­
N.H. Thoan et al. Vacuum 191 (2021) 110389

Fig. 4. (a) PL spectra Sm1/2Na1/2TiO3-modified Bi1/2Na1/2TiO3 compounds as a function of Sm1/2Na1/2TiO3 concentrations, and (b) the comparison of intensity
photoluminescence of 4G5/2 to 6H7/2 transition as a function of Sm1/2Na1/2TiO3 amounts.

Fig. 5. Magnifications of the PL spectra of Sm1/2Na1/2TiO3-modified Bi1/2Na1/2TiO3 compounds as a function of Sm1/2Na1/2TiO3 concentrations in the wavelength
ranges of (a) 550–750 nm, (b) 750–1000 nm, (c) 1000–1600 nm, and (d) 1600–2000 nm.

6
2Na1/2TiO3 and Sm1/2Na1/2TiO3-modified Bi1/2Na1/2TiO3 materials in F3/2, and 6F5/2 ones, respectively. Fig. 5 (c) presented PL spectra in the
different wavelength ranges. Fig. 5 (a) showed the PL spectra in the 1000–1600 nm range, showing four medium peaks. The peak at 1132
wavelength range from 550 to 750 nm. There are four main PL peaks nm might be assigned to the 6H11/2 → 6H7/2 transition. The emission
located at approximately 565, 601, 647, and 711 nm, which are related around 1205 nm arises by 4G5/2 → 6F9/2 or 6F9/2 → 6H7/2 transition. The
to transitions from Sm3+ 4G5/2 level to 6H5/2, 6H7/2, 6H9/2, and 6H11/2 last one at about 1435 nm was assigned to 6F7/2 → 6H7/2 or/and 6F9/2 →
6
ones. Our results were well consistent with the recent observation in PL H9/2 transitions. In Fig. 5 (d), three weak PL peaks at around 1690,
spectra of Sm doped lead-free ferroelectric Bi1/2Na1/2TiO3 materials 1810, and 1950 nm were observed, which might be related to transitions
[32]. Interestingly, we obtained the emission of 4f-4f electron transi­ from the (6H15/2, 6F3/2, 6F1/2) manifold or other 6F levels to 6H levels.
tions of Sm3+ cations randomly incorporated into the host Bi1/2Na1/2 Recently, Starecki et al. found two PL emission transitions by Sm3+ at
TiO3 lattices at near-infrared and infrared regions. In Fig. 5 (b), we around 1650 nm and 1950 nm by 1450 mn excitation, that ascribed to
6
observed three weak PL peaks at around 780, 885, and 959 nm, that F1/2 → 6H5/2 and 6H13/2 → 6H5/2 transitions [74]. Other transitions
were assigned to electron transitions from Sm3+ 4G5/2 level to 6H13/2, related to (6H15/2, 6F3/2, 6F1/2) manifold did not occur in Sm3+-doped

6
­
N.H. Thoan et al. Vacuum 191 (2021) 110389

Ga5Ge20Sb10Se65 fiber because the Sm3+ concentration was low (500 [11] D.D. Dung, N.B. Doan, N.Q. Dung, L.H. Bac, N.H. Linh, L.T.H. Thanh, D.V. Thiet, N.
N. Trung, N.C. Khang, T.V. Trung, N.V. Duc, Role of Co dopants on the structural,
ppm) [60]. In this study, we suggested that at higher Sm3+ concentration
optical and magnetic properties of lead-free ferroelectric Na0.5Bi0.5TiO3 materials,
levels up to 9 mol.% in our studied samples, up-conversion processes J. Sci.: Adv. Mater. Dev. 4 (2019) 584–590, https://doi.org/10.1016/j.
might occur because of interactions between neighboring Sm3+ ions. We jsamd.2019.08.007.
also observed a new peak at 1800 nm. We suggested that the peak was [12] R. Cheng, Z. Xu, R. Chu, J. Hao, J. Du, W. Ji, G. Li, Large piezoelectric effect in Bi1/
2Na1/2TiO3-based lead-free piezoceramics, Ceram. Int. 41 (2015) 8119–8127,
ascribed to 6F3/2 → 6H7/2 transition. However, the evidence is unclear, https://doi.org/10.1016/j.ceramint.2015.03.015.
and one needs further measurements to confirm the assumption. [13] K. Makisumi, Y. Kitanaka, Y. Noguchi, M. Miyayama, Enhanced polarization
properties of ferroelectric (Bi1/2Na1/2)TiO3–Ba(Mg1/3Nb2/3)O3 single crystals
grown under high-pressure oxygen atmosphere, J. Ceram. Soc. Japan 125 (2017)
4. Conclusion 463–467, https://doi.org/10.2109/jcersj2.16285.
[14] Z. Yang, B. Liu, L. Wei, Y. Hou, Structure and electrical properties of (1-x)
The Sm1/2Na1/2TiO3-modified Bi1/2Na1/2TiO3 materials were well Bi0.5Na0.5TiO3-xBi0.5K0.5TiO3 ceramics near morphotropic phase boundary, Mater.
Res. Bull. 43 (2008) 81–89, https://doi.org/10.1016/j.materresbull.2007.02.016.
synthesized by the simple chemical route. The Sm1/2Na1/2TiO3-modified [15] J. Shin, H. Fan, X. Liu, A.J. Bell, Large electrostrictive strain in (Bi0.5Na0.5)TiO3-
Bi1/2Na1/2TiO3 materials were well synthesized by a simple chemical BaTiO3-(Sr0.7Bi0.2)TiO3 solid solution, J. American Ceram. Soc. 97 (2014)
route method. The Sm1/2Na1/2TiO3-modified Bi1/2Na1/2TiO3 com­ 848–853, https://doi.org/10.1111/jace.12712.
[16] L. Liu, H. Fan, Influence of sintering temperatures on the electrical properties of
pounds exhibited the rhombohedral structure that followed the struc­ bismuth sodium titanate based piezoelectric ceramics, J. Electroceram. 16 (2006)
ture of the host Bi1/2Na1/2TiO3 lattice. The optical bandgap of Bi1/2Na1/ 293–296, https://doi.org/10.1007/s10832-006-9867-3.
[17] L. Liu, H. Fan, S. Ke, X. Chen, Effect of sintering temperature on the structure and
2TiO3 materials was slightly varied in the range of 3.07–2.94 eV as a
properties of cerium-doped 0.94(Bi0.5Na0.5)TiO3-0.06BaTiO3 piezoelectric
function of Sm1/2Na1/2TiO3 amounts solid solute into the host Bi1/2Na1/ ceramics, J. Alloys Compd. 458 (2008) 504–508, https://doi.org/10.1016/j.
2TiO3 materials. The strong photoluminescence of Sm1/2Na1/2TiO3- jallcom.2007.04.037.
modified Bi1/2Na1/2TiO3 materials was obtained with the highest in­ [18] D.D. Dung, N.T. Hung, D. Odkhuu, Structure, optical and magnetic properties of
new Bi0.5Na0.5TiO3-SrMnO3− δ solid solution materials, Sci. Rep. 9 (2019) 18186,
tensity for the BNT-5Sm sample containing 5 mol.% Sm1/2Na1/2TiO3 https://doi.org/10.1038/s41598-019-54172-4.
dopants. The Sm1/2Na1/2TiO3-modified Bi1/2Na1/2TiO3 materials [19] M.M. Hue, N.Q. Dung, N.N. Trung, L.H. Bac, L.T.K. Phuong, N.V. Duc, D.D. Dung,
exhibited the NIR region emissions of Sm ions. We expected our work to Tunable magnetic properties of Bi0.5Na0.5TiO3 materials via solid solution of
NiTiO3, Appl. Phys. A 124 (2018) 588, https://doi.org/10.1007/s00339-018-2002-
be an extension of the functional lead-free ferroelectric Bi1/2Na1/2TiO3
x.
materials for optical applications. [20] D.D. Dung, M.M. Hue, N.Q. Dung, N.X. Dung, L.H. Bac, N.N. Trung, N.A. Duc, N.
H. Thoan, Design and characterization of a new (1-x)Na1/2Bi1/2TiO3+xBi(Ti1/2Fe1/
2)O3 solid solution, Vacuum 183 (2021) 109815, https://doi.org/10.1016/j.
Declaration of competing interest vacuum.2020.109815.
[21] D.D. Dung, N.T. Hung, Structural, optical and magnetic properties of (1− x)
Bi0.5Na0.5TiO3+xBaCoO3− δ solid solution systems prepared by the sol–gel method,
The authors declare that they have no known competing financial J. Nanosci. Nanotechnol. 21 (2021) 2604–2612, https://doi.org/10.1166/
interests or personal relationships that could have appeared to influence jnn.2021.19090.
the work reported in this paper. [22] D.D. Dung, N.H. Lam, A.D. Nguyen, N.N. Trung, N.V. Duc, N.T. Hung, Y.S. Kim,
D. Odkhuu, Experimental and theoretical studies on induced ferromagnetism of
new (1-x)Na0.5Bi0.5TiO3+xBaFeO3-δ solid solution, Sci. Rep. 11 (2021) 8908,
Acknowledgments https://doi.org/10.1038/s41598-021-88377-3.
[23] Q. Yao, F. Wang, F. Xu, C.M. Leung, T. Wang, Y. Tang, X. Ye, Y. Xie, D. Sun, W. Shi,
Electric field-induced giant strain and photoluminescence-enhancement effect in
This research is funded by Vietnam National Foundation for Science rare-earth modified lead-free piezoelectric ceramics, ACS Appl. Mater. Interfaces 7
and Technology Development (NAFOSTED) under grant number (2015) 5066–5075, https://doi.org/10.1021/acsami.5b00420.
103.02–2019.366 and this research was partially supported by the [24] H.L. Sun, X. Wu, T.H. Chung, K.W. Kwok, In-situ electric field-induced modulation
of photoluminescence in Pr-doped Ba0.85Ca0.15Ti0.90Zr0.10O3 lead-free ceramics,
Nippon Sheet Glass Foundation.
Sci. Rep. 6 (2016) 28677, https://doi.org/10.1038/srep28677.
[25] H. Sun, X. Wu, D.F. Peng, K.W. Kwok, Room-temperature large and reversible
References modulation of photoluminescence by in situ electric field in ergodic relaxor
ferroelectrics, ACS Appl. Mater. Interfaces 9 (2017) 34042–34049, https://doi.org/
10.1021/acsami.7b09354.
[1] N.D. Quan, L.H. Bac, D.V. Thiet, V.N. Hung, D.D. Dung, Current development in
[26] Y. Lin, C.W. Nan, J. Wang, H. He, J. Zhai, L. Jiang, Photoluminescence of nanosized
lead-free Bi0.5(Na,K)0.5TiO3-based piezoelectric materials, Ann. Mater. Sci. Eng.
Na0.5Bi0.5TiO3 synthesized by a sol–gel process, Mater. Lett. 58 (2004) 829–832,
2014 (2015) 365391, https://doi.org/10.1155/2014/365391.
https://doi.org/10.1016/j.matlet.2003.07.025.
[2] Z. Liu, H. Fan, Y. Zhao, G. Dong, Optical and tunable dielectric properties of
[27] D.D. Dung, N.T. Hung, D. Odkhuu, Magnetic and optical properties of new (1 − x)
K0.5Na0.5TiO3-SrTiO3 ceramics, J. American Ceram. Soc. 99 (2016) 146–151,
Bi0.5Na0.5TiO3 + xCaMnO3− δ solid solution materials, Mater. Sci. Eng., B 263
https://doi.org/10.1111/jace.13864.
(2021) 114902, https://doi.org/10.1016/j.mseb.2020.114902.
[3] Z. Liu, H. Fan, B. Peng, Enhancement of optical transparency in Bi2O3-modified
[28] N.T. Hung, L.H. Bac, N.T. Hoang, P.V. Vinh, N.N. Trung, D.D. Dung, Structural,
(K0.5Na0.5)0.9Sr0.1Nb0.9Ti0.1O3 ceramics for electro-optic applications, J. Mater. Sci.
optical, and magnetic properties of SrFeO3-δ-modified Bi0.5Na0.5TiO3 materials,
50 (2015) 7958–7966, https://doi.org/10.1007/s10853-015-9360-y.
Phys. B Condens. Matter 531 (2018) 75–78, https://doi.org/10.1016/j.
[4] N.T. Hung, N.H. Lam, A.D. Nguyen, L.H. Bac, N.N. Trung, D.D. Dung, Y.S. Kim,
physb.2017.12.021.
N. Tsogbadrakh, T. Ochirkhuyag, D. Odkhuu, Intrinsic and tunable ferromagnetism
[29] D.D. Dung, M.M. Hue, N.Q. Dung, N.H. Lam, L.T.K. Phuong, L.H. Bac, N.N. Trung,
in Bi0.5Na0.5TiO3 through CaFeO3-δ modification, Sci. Rep. 10 (2020) 6189,
N.V. Duc, D. Odkhuu, Enhancing room-temperature ferromagnetism in
https://doi.org/10.1038/s41598-020-62889-w.
Bi0.5Na0.5TiO3 via FeTiO3 solid solution, J. Electroceram. 44 (2020) 129–135,
[5] G.A. Smolenski, V.A. Isupov, A.I. Aganovskaya, New ferroelectrics of complex
https://doi.org/10.1007/s10832-020-00203-w.
composition IV, J. Sov. Phys. Solid State 2 (1961) 2651.
[30] L. Li, M. Xu, Q. Zhang, P. Chen, N. Wang, D. Xiong, B. Peng, L. Liu, Electrocaloric
[6] T. Takennaka, K. Maruyama, K. Sakata, (Bi1/2Na1/2)TiO3-BaTiO3 system for lead-
effect in La-doped BNT-6BT relaxor ferroelectric ceramics, Ceram. Int. 44 (2018)
free piezoelectric ceramics, Jpn. J. Appl. Phys. 30 (1991) 2236, https://doi.org/
343–350, https://doi.org/10.1016/j.ceramint.2017.09.179.
10.1143/JJAP.30.2236.
[31] F. Han, J. Deng, X. Liu, T. Yan, S. Ren, X. Ma, S. Liu, B. Peng, L. Liu, High-
[7] A. Herabut, A. Safari, Processing and electromechanical properties of (Bi0.5Na0.5)(1-
temperature dielectric and relaxation behavior of Yb-doped Bi0.5Na0.5TiO3
1.5x)LaxTiO3 ceramics, J. American Ceram. Soc. 80 (1997) 2954–2958, https://doi.
ceramics, Ceram. Int. 43 (2017) 5564–5573, https://doi.org/10.1016/j.
org/10.1111/j.1151-2916.1997.tb03219.x.
ceramint.2017.01.086.
[8] J.U. Rahman, A. Hussain, A. Magbool, T.K. Song, W.J. Kim, S.S. Kim, M.H. Kim,
[32] T. Wei, F.C. Sun, C.Z. Zhao, C.P. Li, M. Yang, Y.Q. Wang, Photoluminescence
Dielectric, ferroelectric and field-induced strain response of lead-free BaZrO3-
properties in Sm doped Bi1/2Na1/2TiO3 ferroelectric ceramics, Ceram. Int. 39
modified Bi0.5Na0.5TiO3 ceramics, Curr. Appl. Phys. 14 (2014) 331–336, https://
(2013) 9823–9828, https://doi.org/10.1016/j.ceramint.2013.05.094.
doi.org/10.1016/j.cap.2013.12.009.
[33] T. Wei, Z. Zhang, Q.J. Zhou, D.M. An, Z.P. Li, F.C. Sun, Bright green emission in Ho
[9] L. Ju, C. Shi, L. Sun, Y. Zhang, H. Qin, J. Hu, Room-temperature magnetoelectric
doped Bi1/2Na1/2TiO3 ferroelectric ceramics, Mater. Lett. 115 (2014) 129–131,
coupling in nanocrystalline Na0.5Bi0.5TiO3, J. Appl. Phys. 116 (2014), 083909,
https://doi.org/10.1016/j.matlet.2013.10.051.
https://doi.org/10.1063/1.4893720.
[34] P. Du, L. Luo, W. Li, Y. Zhang, H. Chen, Electrical and luminescence properties of
[10] L.T.H. Thanh, N.H. Tuan, L.H. Bac, D.D. Dung, P.Q. Bao, Influence of fabrication
Er-doped Bi0.5Na0.5TiO3 ceramics, Mater. Sci. Eng., B 178 (2013) 1219–1223,
condition on the microstructural and optical properties of lead-free ferroelectric
https://doi.org/10.1016/j.mseb.2013.08.007.
Bi0.5Na0.5TiO3 materials, Commun. Phys. 26 (2016) 51–57, https://doi.org/
10.15625/0868-3166/26/1/7354.

7
N.H. Thoan et al. Vacuum 191 (2021) 110389

[35] M. Dunce, G. Krieke, E. Birks, L. Bikse, M. Antonova, A. Sarakovkis, The role of [58] D.D. Dung, N.T. Hung, Magnetic properties of (1-x)Bi0.5Na0.5TiO3+xSrCoO3-δ solid-
structural disorder on luminescence of Eu-doped Na0.5Bi0.5TiO3, J. Appl. Phys. 128 solution materials, Appl. Phys. A 126 (2020) 240, https://doi.org/10.1007/
(2020) 244104, https://doi.org/10.1063/5.0031305. s00339-020-3409-8.
[36] K.R. Kandula, S. Asthana, S.S.K. Raavi, Multifunctional Nd3+ substituted [59] C.M. Reddy, B.D.P. Raju, N.J. Sushma, N.S. Dhoble, S.J. Dhoble, A review on
Na0.5Bi0.5TiO3 as lead-free ceramics with enhanced luminescence, ferroelectric and optical and photoluminescence studies of RE3+ (RE=Sm, Dy, Eu, Tb and Nd) ions
energy harvesting properties, RSC Adv. 8 (2018) 15282–15289, https://doi.org/ doped LCZSFB glasses, Renew. Sustain. Energy Rev. 51 (2015) 566–584, https://
10.1039/C8RA01349G. doi.org/10.1016/j.rser.2015.06.025.
[37] H. Sun, D. Peng, X. Wang, M. Tang, Q. Zhang, X. Yao, Strong red emission in Pr [60] D.D. Dung, N.B. Doan, N.Q. Dung, N.H. Linh, L.H. Bac, L.T.H. Thanh, N.N. Trung,
doped (Bi0.5Na0.5)TiO3 ferroelectric ceramics, J. Appl. Phys. 110 (2011), 016102, N.V. Duc, L.V. Cuong, D.V. Thiet, S. Cho, Tunable magnetism of Na0.5Bi0.5TiO3
https://doi.org/10.1063/1.3606425. materials via Fe defects, J. Supercond. Nov. Magnetism 32 (2019) 3011–3018,
[38] S.K. Barik, R.N.P. Choudhary, P.K. Mahapatra, Impedance spectroscopy study of https://doi.org/10.1007/s10948-019-05163-z.
Na1/2Sm1/2TiO3 ceramic, Appl. Phys. A 88 (2007) 217–222, https://doi.org/ [61] L.T.H. Thanh, N.B. Doan, L.H. Bac, D.V. Thiet, S. Cho, P.Q. Bao, D.D. Dung, Making
10.1007/s00339-007-3990-0. room-temperature ferromagnetism in lead-free ferroelectric Bi0.5Na0.5TiO3
[39] F. Luo, B. Tang, Y. Yuan, Z. Fang, S. Zhang, Microstructure and microwave material, Mater. Lett. 186 (2017) 239–242, https://doi.org/10.1016/j.
dielectric properties of Na1/2Sm1/2TiO3 filled PTFE, an environmental friendly matlet.2016.09.105.
composites, Appl. Sur. Sci. 436 (2018) 900–906, https://doi.org/10.1016/j. [62] D.D. Dung, N.Q. Dung, N.B. Doan, N.H. Linh, L.H. Bac, N.N. Trung, N.V. Duc, L.T.
apsusc.2017.11.207. H. Thanh, L.V. Cuong, D.V. Thiet, S. Cho, Defect-mediated room temperature
[40] Z.X. Fang, B. Tang, E. Li, S.R. Zhang, High-Q microwave dielectric properties in the ferromagnetism in lead-free ferroelectric Na0.5Bi0.5TiO3 materials, J. Supercond.
Na0.5Sm0.5TiO3 + Cr2O3 ceramics by one synthetic process, J. Alloys Compd. 705 Nov. Magnetism 33 (2020) 911–920, https://doi.org/10.1007/s10948-019-05399-
(2017) 456–461, https://doi.org/10.1016/j.jallcom.2017.01.200. 9.
[41] F. Liu, J. Qu, C. Yuan, G. Chen, P. Qin, X. Huang, Microwave dielectric properties [63] L.T.H. Thanh, N.B. Doan, N.Q. Dung, L.V. Cuong, L.H. Bac, N.A. Duc, P.Q. Bao, D.
of Na0.5Sm0.5TiO3-based ceramics, J. Mater. Sci. Mater. Electron. 28 (2017) D. Dung, Origin of room temperature ferromagnetism in Cr-doped lead-free
3052–3059, https://doi.org/10.1007/s10854-016-5892-4. ferroelectric Bi0.5Na0.5TiO3 materials, J. Electron. Mater. 46 (2017) 3367–3372,
[42] R.D. Shannon, Revised effective ionic radii and systematic studies of interatomic https://doi.org/10.1007/s11664-016-5248-0.
distances in halides and chalcogenides, Acta Crystallogr. A 32 (1976) 751–767, [64] N.H. Linh, N.H. Tuan, D.D. Dung, P.Q. Bao, B.T. Cong, L.T.H. Thanh, Alkali metal-
https://doi.org/10.1107/S0567739476001551. substituted-based perovskite compounds: a DFT study, J. Sci.: Adv. Mater. Dev. 4
[43] M. Uichiro, Hume-rothery Rules for Structurally Complex Alloy Phases, Taylor & (2019) 492–498, https://doi.org/10.1016/j.jsamd.2019.06.005.
Francis, 2010. [65] M. Zeng, S.W. Or, H.L.W. Chan, First-principle study on the electronic and optical
[44] W. Hume-Rothery, Atomic Theory for Students of Metallurgy, The Institute of properties of Na0.5Bi0.5TiO3 lead-free piezoelectric crystal, J. Appl. Phys. 107
Metals, London, 1969. (2010), 043513, https://doi.org/10.1063/1.3309407.
[45] C.B. Carter, M.G. Norton, Ceramic Materials: Science and Engineering, Springer, [66] N. Pradhani, P.K. Mahapatra, R.N.P. Choudhary, Effect of cerium oxide addition on
2007. optical, electrical and dielectric characteristics of (Bi0.5Na0.5)TiO3 ceramics,
[46] Y.S. Sung, J.M. Kim, J.H. Cho, T.K. Song, M.H. Kim, H.H. Chong, T.G. Park, D. Do, J. Phys.: Mater. 1 (2018), 015007, https://doi.org/10.1088/2515-7639/aacff0.
S.S.S. Kim, Effects of Na nonstoichiometry in (Bi0.5Na0.5+x)TiO3 ceramics, Appl. [67] K.R. Kandula, S. Asthana, S.S.K. Raavi, Multifunctional Nd3+ substituted
Phys. Lett. 96 (2010), 022901, https://doi.org/10.1063/1.3275704. Na0.5Bi0.5TiO3 as lead-free ceramics with enhanced luminescence, ferroelectric and
[47] M. Spreitzer, M. Valant, D. Suvorov, Sodium deficiency in Na0.5Bi0.5TiO3, J. Mater. energy harvesting properties, RSC Adv. 8 (2018) 15282–15289, https://doi.org/
Chem. 17 (2007) 185–192, https://doi.org/10.1039/B609606A. 10.1039/C8RA01349G.
[48] I.K. Jeong, Y.S. Sung, T.K. Song, M.H. Kim, A. Llobet, Structural evolution of [68] Y. An, C. He, C. Deng, Z. Chen, H. Chen, T. Wu, Y. Lu, X. Gu, J. Wang, Y. Liu, Z. Li,
bismuth sodium titanate induced by A-site non-stoichiometry: neutron powder Poling effect on optical and dielectric properties of Pr3+-doped Na0.5Bi0.5TiO3
diffraction studies, J. Kor. Phys. Soc. 67 (2015) 1583–1587, https://doi.org/ ferroelectric single crystal, Ceram. Int. 46 (2020) 4664–4669, https://doi.org/
10.3938/jkps.67.1583. 10.1016/j.ceramint.2019.10.197.
[49] L. Luo, P. Du, W. Li, W. Tao, H. Chen, Effects of Er doping site and concentration on [69] M. Benyoussef, H. Zaari, J. Belhadi, Y.E. Amraoui, H.E. Zahraouy, A. Lahmar, M.
piezoelectric, ferroelectric and optical properties of ferroelectric Na0.5Bi0.5TiO3, E. Marssi, Effect of rare earth on physical properties of Na0.5Bi0.5TiO3 system: a
J. App. Phys. 114 (2013) 124104, https://doi.org/10.1063/1.4823812. density functional theory investigation, J. Rare Earths (2020), https://doi.org/
[50] S. Behara, R.H. Krishna, M. Murlidhar, M. Murakami, M. Irfan, S. Najma, 10.1016/j.jre.2020.12.002.
T. Thomas, Amphotericity-spectroscopy correlation in Eu doped sodium bismuth [70] S. Behara, G.S. Priyanga, T. Thomas, Strain-induced effects in the electronic and
titanate (Na0.5Bi0.5TiO3), Materialia 7 (2019) 100426, https://doi.org/10.1016/j. optical properties of Na0.5Bi0.5TiO3: an ab-initio study, Mater. Today Commun. 24
mtla.2019.100426. (2020) 101348, https://doi.org/10.1016/j.mtcomm.2020.101348.
[51] C. Chatzichristodoulou, P. Norby, P.V. Hendriksen, M.B. Mogensen, Size of oxide [71] A. Herrera, A. Becerra, N.M. Balzaretti, Novel NIR emission 4G5/2 to 6F11/2 and
vacancies in fluorite and perovskite structured oxides, J. Electroceram. 34 (2015) efficient multichannel emissions of Sm3+ doped GeO2-PbO glass, J. Lumin. 188
100–107, https://doi.org/10.1007/s10832-014-9916-2. (2017) 193–198, https://doi.org/10.1016/j.jlumin.2017.04.033.
[52] X. Liu, H. Fan, J. Shi, L. Wang, H. Du, Enhanced ionic conductivity of Ag addition [72] A. Herrera, R.G. Fernandes, A.S.S. de Camargo, A.C. Hernandes, S. Buchner,
in acceptor-doped Bi0.5Na0.5TiO3 ferroelectrics, RSC Adv. 6 (2016) 30623–30627, C. Jacinto, N.M. Balzaretti, Visible–NIR emission and structural properties of Sm3+
https://doi.org/10.1039/C6RA00682E. doped heavy-metal oxide glass with composition B2O3–PbO–Bi2O3–GeO2,
[53] B.K. Barick, K.K. Mishra, A.K. Arora, R.N.P. Choudhary, D.K. Pradhan, Impedance J. Lumin. 171 (2016) 106–111, https://doi.org/10.1016/j.jlumin.2015.10.065.
and Raman spectroscopic studies of (Na0.5Bi0.5)TiO3, J. Phys. D Appl. Phys. 44 [73] Q. Sheng, Y. Shen, S. Liu, W. Li, D. Chen, Blue-white tunable luminescence for
(2011) 355402, https://doi.org/10.1088/0022-3727/44/35/355402. white light-emitting diodes and wideband near-infrared luminescence from Sm3+-
[54] M.K. Niranjan, T. Karthik, S. Asthana, J. Pan, U.V. Waghmare, Theoretical and doped borophosphate glass, Appl. Phys. Lett. 101 (2012), 061904, https://doi.org/
experimental investigation of Raman modes, ferroelectric and dielectric properties 10.1063/1.4743008.
of relaxor Na0.5Bi0.5TiO3, J. Appl. Phys. 113 (2013) 194106, https://doi.org/ [74] F. Starecki, A. Braud, N. Abdellaoui, C. Boussard-plédel, B. Bureau, V. Nazabal, 7 to
10.1063/1.4804940. 8 μm emission from Sm3+ doped selenide fibers, Opt Express 26 (2018)
[55] L. Liu, M. Knapp, H. Ehrenberg, L. Fang, H. Fan, L.A. Schmitt, H. Fuess, M. Hoelzel, 26462–26469, https://doi.org/10.1364/OE.26.026462.
H. Dammak, M.P. Thi, M. Hinterstein, Average vs. local structure and composition- [75] J. He, J. Zhang, H. Xing, H. Pan, X. Jia, J. Wang, P. Zheng, Thermal stable
property phase diagram of K0.5Na0.5NbO3-Bi1/2Na1/2TiO3 system, J. Eur. Ceram. ferroelectricity and photoluminescence in Sm-doped 0.8(Bi0.5Na0.5)TiO3-
Soc. 37 (2017) 1387–1399, https://doi.org/10.1016/j.jeurceramsoc.2016.11.024. 0.25SrTiO3 ferroelectric ceramics, Ceram. Int. 43 (2017) 250–255, https://doi.org/
[56] W. Liu, X. Ma, S. Ren, X. Lei, L. Liu, Tunable phase transition in 10.1016/j.ceramint.2016.09.146.
(Bi0.5Na0.5)0.94Ba0.06TiO3 by B-site cations, Appl. Phys. A 126 (2020) 269, https:// [76] J.C.C.A. Diaz, M. Venet, F. Cordero, J.P.S. d Silva, Anelastic and optical properties
doi.org/10.1007/s00339-020-3448-1. of Bi0.5Na0.5TiO3 and (Bi0.5Na0.5)0.94Ba0.06TiO3 lead-free ceramic systems doped
[57] X. He, S. Zhou, Z. Yuan, Y. Li, Z. Lan, K. Chen, F. Han, X. Lei, L. Liu, The origin of with donor Sm3+, J. Alloys Compd. 746 (2018) 648–652, https://doi.org/10.1016/
weak coupling between polar clusters in Ta2O5-doped 0.94Bi0.5Na0.5TiO3- j.jallcom.2018.02.303.
0.06BaTiO3, Europhys. Lett. 129 (2020) 37004, https://doi.org/10.1209/0295- [77] I.P. Rosales, R.L. Juarez, G.L. Pacheco, C. Falcony, F. Gonzalez, Near infrared
5075/129/37004. photon-downshifting in Yb3+-doped titanates: the influence of intrinsic defects,
J. Alloys Compd. 834 (2020) 155081, https://doi.org/10.1016/j.
jallcom.2020.155081.

You might also like