Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Inorganic Chemistry Communications 146 (2022) 110119

Contents lists available at ScienceDirect

Inorganic Chemistry Communications


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

Short communication

Structural, morphological, optical and dielectric properties of sodium


bismuth titanate ceramics
Najah Rhimi a, *, N. Dhahri a, b, M. Khelifi a, E.K. Hlil c, J. Dhahri a
a
Laboratoire de la Matière Condensée et des Nanosciences, Faculté Des Sciences de Monastir, Université de Monastir, LR11ES40, 5000 Monastir, Tunisia
b
Physics Department, Faculty of Science, Northern Border University, Arar, Saudi Arabia
c
Institut Néel, CNRS et Université Joseph Fourier, BP 166, F-38042 Grenoble cedex 9, France

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

Keywords: This work focuses on the structural, optical and electrical properties of sodium bismuth titanate Bi1/2Na1/2TiO3
Perovskite (NBT) lead-free ceramic for its possible application in capacitors, batteries and as an electrolyte in SOFC. NBT
X-ray diffraction material was prepared by the solid-state method. X-ray diffraction analysis performed at room temperature
Reflectance
revealed a rhombohedral structure with R3c space group. Rietveld refinement confirmed a good agreement
Dielectric
between the calculated and the observed pattern. Scanning electron microscopy analysis revealed a poly­
Conductivity
crystalline nature of the material with a grain size of 29.17 µm. FT-IR analysis showed the presence of different
bonds. The band gap energy was extracted through diffuse reflectance spectroscopy and was found to be 2.61 eV.
Detailed studies of dielectric and impedance properties carried out in the frequency range of 40–107 Hz at
different temperatures (from 400 K to 700 K) provided many interesting properties. The dielectric loss and
dielectric constant with frequency increase. This behavior could be interpreted by the Maxwell–Wagner type of
polarization in agreement with Koop’s theory. To explain the complex impedance plane plots, an equivalent
circuit template was employed, reflecting semiconducting behavior of NBT. The Ac-conductivity spectrum
satisfied Jonscher’s power law below 460 K and followed Jonscher’s double power law (DPL) above 480 K. The
correlated barrier hopping (CBH) and small polaron hopping (SPH) models could explain the conduction
mechanism at the studied temperature range. The study of the Dc-electrical conductivity confirmed that NBT
followed the variable range hopping (VRH) model. The present work could give further insights into this kind of
material from a physical point of view and thus enhance its presence in the industrial field.

1. Introduction several lead-free materials, sodium bismuth titanate Bi1/2Na1/2TiO3


(NBT), could be a good candidate for lead-free ceramics in view of its
Various attempts have recently been made to study perovskite high Curie temperature (Tc = 320 ◦ C) and high remanent polarization
structure (ABO3) based oxide ceramics due to their variety of properties (Pr = 38 µC/cm2) at room temperature [20–21]. Moreover, Pb2+ is
including ferroelectric, dielectric, piezoelectric, pyroelectric and optical responsible for the large polarization due to its long pair effect of 6 s
properties. In fact, they could be used in an extensive range of appli­ valence shell electron in PZT ceramics. Bi3+ has the same electronic
cations such as solid oxide fuel cells, actuators, transducers, micro- configuration as Pb2+ (isoelectronic) in PZT systems, which motivated
electromechanical systems, filters, oxygen separation membranes, res­ scientists to consider the NBT system a good substitute for PZT. NBT,
onators, multilayer capacitors, non-volatile ferroelectric random access first reported by Smolenskii et al. [22], is a lead-free perovskite material
memory and batteries [1–16]. Many of the perovskite materials used for having a mixture of two different cations (Na+ and Bi3+) at site A, while
device applications are Pb-based. In particular PZT (PbZrxTi1-xO3), PLZT the oxygen anions form octahedra whose centers are occupied by the
(Pb(1− x)Lax(ZryT1− y)O3) and PMN-PT (Pb(Mg1/3Nb2/3)O3-PbTiO3) Ti4+ cations. According to previous studies [23–25], NBT is a ferro­
[17–19] are very hazardous for the environment despite their excellent electric complex with a rhombohedral symmetry at room temperature.
physical properties. Therefore, the demand for using lead-free materials Furthermore, the temperature in situ works, performed by X-ray,
that are environmentally friendly has been increasing. Among the Neutron diffraction and Raman spectroscopy revealed that at least two

* Corresponding author.
E-mail address: [email protected] (N. Rhimi).

https://doi.org/10.1016/j.inoche.2022.110119
Received 7 June 2022; Received in revised form 26 September 2022; Accepted 16 October 2022
Available online 4 November 2022
1387-7003/© 2022 Elsevier B.V. All rights reserved.
N. Rhimi et al. Inorganic Chemistry Communications 146 (2022) 110119

phase transitions occur [26–27]. Indeed, the rhombohedral phase with (400–700 K).
the polar R3c space group changes to the tetragonal phase with the P4bm
space group above 400 ◦ C. According to Vakhrushev et al. [28], the two 3. Results and discussion
phases coexist at the temperature range (250–400 ◦ C). The cubic phase
(Pm 3 m) appears at about 520 ◦ C. Concerning the dielectric properties 3.1. Structural properties
of NBT, many authors have reported that the evolution of the dielectric
constant with temperature has two anomalies associated with the phase The XRD pattern of NBT ceramic is illustrated in Fig. 1a. As can be
transitions from ferroelectric (FE) to antiferroelectric (AFE) and from seen, a Well-defined splitting of (1 1 0)R, (2 0 2)R, (0 2 4)R and (3 0 0)R
antiferroelectric to paraelectric within NBT [29,30]. The first anomaly is peaks was noticed, which indicates the rhombohedral symmetry with
frequency dependent at around 200 ◦ C, showing the relaxor behavior. small amount of Na0.5Bi4.5Ti4O15 as an impurity phase [35].A repre­
The other anomaly does not depend on the frequency possessing a high sentative selection of the Bragg peaks of NBT is displayed in Fig. 1b. A
dielectric permittivity at about 320 ◦ C. Above 400 ◦ C, the Curie-Weiss small shoulder was observed on the left side of these reflection peaks
law is achieved corresponding to the paraelectric behavior of NBT. (indicated by arrow). According to Xie et al. [36], these shoulders are
NBT-based ceramics can be synthesized by several methods such as often attributed to rhombohedral symmetry. We carried out the struc­
co-precipitation, sol–gel method, molten salt route and solid-state re­ ture refinement with R3C space group in the rhombohedral unit cell by
action. The most advantageous method is the solid-state reaction owing Rietveld analysis (see Fig. 1c). A good agreement between the simulated
to its low cost and simplicity in process and handling. During solid-state and the observed results was noticed. The results of refinement are
reaction, the sintering temperature and sintering time are the most collected in Table 1. For NBT, the lattice parameters obtained experi­
important parameters influencing the properties of the obtained ce­ mentally are in good agreement with those of Jones [37] a = 5.4887 Å
ramics. By varying these parameters, one may expect the variation of the and c = 13.5048 Å.
different dielectric properties of NBT. As reported by Y. Guo et al. [31] To describe the geometric packing within perovskite structure,
the dielectric permittivity and loss reach values of εmax = 2854 and tolerance factor (t) suggested by Goldschmidt is used. This parameter is
Tanδ = 0.0430 respectively, at 100 kHz for NBT sintered at 1150 ◦ C for defined by the following relation [38]:
4 h. B. Parija et al. [32] showed that the dielectric permittivity and loss
〈rA 〉 + 〈rO 〉
obtained for NBT sintered at 1050 ◦ C for 2 h are εmax = 3500 and Tanδ = t = √̅̅̅ (1)
0.0533 at 100 kHz respectively. B.K. Barick et al. [33] also found that 2(〈rB 〉 + rO )
the dielectric constant reaches 2573 for NBT sintered at 1080 for 4 h. Where 〈rA 〉, 〈rB〉 and 〈rO〉 are the ionic radii of A, B and O sites,
In this paper, we have prepared NBT ceramic by solid-state process respectively. In our case, t = 0.9762 (0.96 < t < 1); therefore our
sintered at 1100 ◦ C for 2 h. The structural characterizations were per­ compound belongs to stable perovskite structure.
formed using X-ray Diffraction (XRD), Scanning Electron Microscopy The crystallite size (D) of our compound was calculated from the
(SEM) and Energy-Dispersive X-ray analysis (EDX). Fourier-Transform broadening of X-ray diffraction peaks using two methods. Scherer’s
Infrared spectroscopy (FT-IR) was performed for detecting different formula can be expressed as follows [39]:
functional groups present in the sample and the optical band gap was
examined using Ultraviolet–Visible spectroscopy (UV–vis). In order to DSh =
0.9λ
(2)
understand the conductivity mechanism, the Ac-conductivity of our βcosθ
sample was studied. Where λ is the wavelength of CuKα radiation (λ = 1.5406 Å), θ is the
angle of the most intense peak (1 1 0) and β is the full width at half
2. Experimental details maximum (FWHM). The Williamson–Hall formula is given as follows
[40]:
2.1. Sample preparation

βcosθ = + 4εsinθ (3)
DWH
Na0.5Bi0.5TiO3 was synthesized by the solid-state method. The
starting materials: Na2CO3 (99 %), Bi2O3 (98 %) and TiO2 (99 %) were Where ε is the elastic strain introduced inside the material. Fig. 1d
weighed by stoichiometry and ground for 30 min in ethanol with an shows the variation of β cos θ versus 4sinθ. The grain size estimated by
agate mortar to homogenize the mixture. Then, the mixed powder was Scherrer’s method was smaller than that obtained by Williamson –Hall
calcined for 8 h at 800 ◦ C. The relatively low calcinations temperature is formula. This difference is due to the lattice strain correction term in the
due to the low melting points of Na2CO3 (851 ◦ C) and Bi2O3 (825 ◦ C) calculations.
[34]. The powder obtained was ground again for 30 min and calcined for Fig. 2a shows the SEM morphology for the NBT ceramic. As
a second time at 1100 ◦ C for 2 h. After that it was compressed into discs observed, the ceramic exhibited high-quality with few pores and clear
with a diameter of 8 mm. Finally, the pellets were sintered at 1100 ◦ C for grain boundaries. The inset of Fig. 2a illustrates the histograms of
2 h. average grains size, estimated using ImageJ software. Fig. 2b displays
the EDX spectrum confirming the existence of all elements in the sample.

2.2. Sample characterization


3.2. FT-IR spectroscopy
The phase structure of our material was investigated by the XRD
technique with a PANAlytical X’Pert Pro diffractometre using CuKα FT-IR analysis is a technique used for detecting different functional
radiation (λ = 1.540598 Å). To check the microstructure, a Scanning groups present in the sample and determining the quality of the sample.
Electronic Microscope (SEM) type JEOL JSM.6390L in conjunction with The FT-IR spectra of NBT measured in the wavenumber ranging 400–
the EDS analysis was used. The FT-IR spectrum was obtained with a 4000 cm− 1 is displayed in Fig. 3. The band appearing at ~418 can be
Shimadzu FTIR-8400 spectrophotometer in the wavenumber range attributed to metal–oxygen vibrations [41]. A strong band observed
4000–400 cm− 1 at room temperature. Concerning the reflectance mea­ around ~535 can be assigned to Ti-O vibrations [42]. The bands
surements, an ocean optic spectrometer (2000 USB) in the wavelength observed at ~966 and at ~2935 cm− 1 may be assigned to the C–O
range 300–1000 nm was used at room temperature. For the dielectric stretching vibration of the carboxylate group and to the Ti-O-Ti vibra­
properties, a Wayne-Kerr 6425 component analyzer was used in the tions of TiO6 octahedra. The bands noticed at ~2935 cm− 1 may be
frequency range (40 Hz- 10 MHz) and in the temperature range attributed to and BO6 octahedral groups in perovskite structure

2
N. Rhimi et al. Inorganic Chemistry Communications 146 (2022) 110119

Fig. 1. (a) Room temperature powder XRD. (b) Representative selection of the Bragg peaks. (c) The structure refinement pattern. (d) The variation of βcosθ versus
4sinθ of NBT ceramic.

Table 1 K (1 − R)2
Structural parameters obtained from Rietveld refinement of XRD data for NBT F(R) = = (4)
S 2R
ceramic.
Where, R, K and S are the relative reflectance ratio, the absorption
R3c phase
coefficient and the scattering coefficient, respectively. The value of the
Cell parameters a = b (Å) 5.4831(5) band transition gap Eg of NBT is calculated using the Tauc equation:
c (Å) 13.4879(4)
V (Å3) 351.1735 αhυ = A(hυ − Eg)n (5)
Atomic positions Na/Bi (x,y,z) (0,0,0.25)
Ti (x,y,z) (0,0,0) Where h is Planck’s constant, υ is the incident light frequency and A
O (x,y,z) (0.4598(1),0,0.25) is a material-dependent constant. Zeng et al. [44] argue that the NBT
2
Biso (Å ) Na/Bi 1.7046(9)
optical properties are governed by charge transfer transitions between
Ti 0.7780(1)
O 0.6792(1)
oxygen 2p and titanium 3d. n = 1/2 and 2 for a direct allowed transition
R-factors (%) Rexp 6.39 and for an indirect allowed transition, respectively. The band gap energy
Rp 6.33 value (Eg = 2.61 eV) can be calculated from the extrapolation of the
Rwp 8.39 linear part of the (αhν)2 vs energy (hν) plot (inset of Fig. 4).
χ2 1.73

3.4. Impedance analysis


providing BO bonds along c-axis respectively [41].
The impedance spectroscopy (IS) is an experimental technique used
3.3. UV–visible spectroscopy to understand the electrical properties of the polycrystalline sample in a
broad range of frequency from 40 Hz to 10 MHz at different tempera­
Following the FT-IR measurements, the diffuse reflectance spectra of tures (from 400 K to 700 K). The complex impedance Z* obtained from
our sample was tested. Fig. 4 illustrates the diffuse reflectance plot of this technique can be decomposed into real (Z′ ) and imaginary Z′′ part. It
NBT ceramic at room temperature. The conversion of the reflectance to is expressed as follows:
absorbance data is obtained according to the Kubelka–Munk equation
(K-M) [43]: Z* = Z’ + jZ’’ (6)
Fig. 5a displays the frequency dependence of Z′ as a function of
frequency at different temperatures for NBT. At low frequencies, Z′

3
N. Rhimi et al. Inorganic Chemistry Communications 146 (2022) 110119

Fig. 4. Diffused reflectance spectrum of NBT ceramic.

Fig. 2. (a) SEM image, inset histogram of average grains size. (b) EDX analysis
of the element distribution for the NBT ceramic.

Fig. 3. FT-IR spectra of NBT ceramic.

Fig. 5. (a)The real part of impedance (Z′ ) vs Frequency. (b) Imaginary part of
decreased with temperature increase suggesting a negative temperature
impedance (Z′ ′ ) vs frequency for NBT ceramic.
coefficient of resistance (NTCR) usually observed in semiconductors
[45]. This behaviour is substantially due to a decrease in the density of
impedance (Z′′ ) value at different temperatures. As can be noticed, Z′′
trapped charges and an increase in the mobility of charge carriers [45].
reached a peak for all temperatures, suggesting the appearance of
At high frequencies, Z′ does not depend on temperature because of the
electrical relaxation. When temperature increase, Z′′ intensity decreased
liberation of space charges owing to the drop of the barrier height,
and shifted towards the higher frequency side, indicating the tempera­
leading to an improved conductivity with the reduction in impedance
ture dependence of the relaxation process, which induced a relaxation
[46].
time spreading [47,48]. Furthermore, the asymmetric broadening of the
Fig. 5b depicts the frequency dependence of the imaginary

4
N. Rhimi et al. Inorganic Chemistry Communications 146 (2022) 110119

Z’’ peak indicates the combined effect of grains and grain boundaries. In Fig. 6.b the imaginary part of electric modulus is represented as a
The peaks due to grain boundaries and grain cannot be differentiated function of frequency at different temperatures for NBT. Below 460 K,
because of the small difference in their time constants [49]. Another two peaks were noticed in M′′ , which confirms the coexistence of both
stronger peak appeared in the low frequency side, which can be asso­ grain and grain boundary effect in the material. However, above that
ciated to electrode-semiconductor contact. temperature, only a single relaxation peak was observed.

3.6. Complex impedance spectra


3.5. Modulus analysis
The Nyquist plot between Z′ and Z′′ at different temperatures is
Complex modulus spectroscopic study is another approach to displayed in Fig. 7a. Two depressed semi-circular arcs were noticed. At
examine the electrical properties of compounds. It suppresses the elec­ the high frequency region, the depressed semi-circular arc can be
trode polarization effects and magnifies those effects which cannot be attributed to grains and grain boundaries conduction effects and at the
distinguished from impedance analysis. It can be introduced as the low frequency side it can be attributed to electrode/interfacial effects.
reciprocal of complex dielectric permittivity: As temperature increased, the radii of the semicircles decreased, indi­
1 cating the decrease of the resistivity of the sample [51]. The experi­
M* = = M’ + jM’’ (7)
ε* mental data were fitted (via Z-view software) and the model proposed is
a series array of parallel combination of (Rg-CPE1), (Rgb-CPE2) and (re-
Where M’ is the real part and M’’ is the imaginary part of electric
CPE3) as shown in Fig. 7b. Here CPE represents the constant phase
modulus.
element and Rg, Rgb and Re are the grain resistance, the grain boundary
The frequency dependence of the real part of modulus (M′ ) at
resistance and electrode resistance, respectively. A close agreement was
different temperatures is shown in Fig. 6a. As can be seen, The M′ value
found between the fitted and observed values, which suggest the cor­
decreased as temperature increased, proving that the conduction can be
rectness of choosing the equivalent circuit. The fitted Rg and Rgb pa­
due to the short range mobility of charge carriers [50]. At low fre­
rameters are plotted in Fig. 7c. Rg and Rgb value decreased with
quencies, a frequency independent plateau appeared and M′ value ten­
temperature increase, which is a typical behavior of semiconductors.
ded toward zero at all temperatures. In addition to the frequency
Fig. 7d shows the combined diagram of the imaginary modulus (M′′ )
increase, a continuous dispersion was noticed and reached a maximum
and impedance (Z′′ ) as a function of frequency for 600 K. The maxima of
value (M∞ ) = (ε1∞ ) due to the relaxation process at high frequencies.
the imaginary modulus (M′′ ) and impedance (Z′′ ) are not frequency
coincident, indicating the deviation from ideal Debye behavior. This
mismatch of two peaks suggests also that the conduction process is
localized [52].

3.7. Dielectric analysis

To study the dielectric properties of our sample, the complex


permittivity is defined from the complex impedance as [53]:

1 Z’’
ε* = *
= 2
= ε’ + jε’’ (8)
jωC0 Z ωC0 (Z’ + Z’’2 )
Where ε′ and ε′′ are the real and imaginary parts of the permittivity,
respectively, and C0 is the geometrical capacity.
The variation of ε′ as a function of frequency (from 40 Hz to 107 Hz)
at different temperatures (from 400 K to 700 K) is depicted in Fig. 8a. In
the low frequency region, the έ value decreased rapidly, while in the
high frequency region, this decrease in values is almost constant. This
behaviour can be explained by Maxwell–Wagner polarization theory
and Koop’s phenomenological theory [54,55]. In Accordance with this
model, a dielectric medium is supposed to be composed of highly con­
ducting grains which are more effective at lower frequencies and poorly
conducting grain boundary appeared to be more effective at higher
frequencies. The high value of ε′ at lower frequencies can be attributed
to interfacial polarization where the charge carriers were accumulated
at the grain boundaries [56].
To promote our work, we must compare our results with similar
studies. At TC and at 100 kHz, the dielectric constant reached a value of
3871, which is lower than that of Y. Guo et al. (εmax = 2854) [31], B.
Parija et al. (εmax = 3500) [32] and than that of B.K. Barick et al.(εmax =
2573) [33].
The dielectric loss (tanδ) gives information about the energy loss. It
can be estimated using the following relation:
ε’’
tanδ = (9)
ε’
Fig. 8b displays the frequency variation of tanδ at different temper­
Fig. 6. (a)The real part of modulus (M′ ) vs Frequency. (b) Imaginary part of atures. A single loss peak was observed. This behavior can be described
modulus (M′ ′ ) vs frequency for NBT ceramic. by Rezlescu model [57]. According to this model, this peak occurs when

5
N. Rhimi et al. Inorganic Chemistry Communications 146 (2022) 110119

Fig. 7. (a) Nyquist plot from 400 to 700 K. (b) fitted data at 500 K and the model proposed. (c) Variation of Rg and Rgb with temperature. (d) Frequency dependent
combined Z′ ′ and M′ ′ plot at 600 K for NBT ceramic.

the frequency of charge hopping between Ti3+ and Ti4+ matches with into two sub-ranges, from 400 to 460 K and from 480 to 700 K. Hence,
that of the applied field. the Ac-conductivity at the first temperature range (400–460 K) is
expressed by Jonscher’s power law [60]:

3.8. Conductivity analysis σ Ac = σ Dc + A × ωn (12)


However, the frequency dependence of the Ac-conductivity in the
Electrical conductivity is a powerful tool used to study the nature of
second region (480–700 K) follows Jonscher’s double power law (DPL):
charge carriers and transport processes. It may be calculated by the
following expression: σ Ac (ω) = σDc + A × ωn1 + B × ωn2 (13)
G×e Where σDc is the Dc-conductivity, ω is an angular frequency, A is a
σ= (10)
S temperature dependent parameter giving information about the
Where G is the conductance, S and e are the thickness and the area of strength of the polarizability (non-ideal conductivity) and n is the power
the pellet, respectively. law exponent representing the degree of interaction between mobile
The total conductivity (σ) is given by following equation: charge carriers and their environments (0 < n < 1). These interactions
can be either translational or localized in nature [61]. Fig. 9b illustrates
σ = σ Dc (T) + σAc (ω, T) (11) an example of fitting the experimental data for T = 600 K.
Where σ Dc is associated to the drift mobility of free charge carriers To explain the conduction mechanism, several theoretical models
and varies with temperature but does not show any variation with fre­ have been proposed based on the temperature behavior of the exponent
quency and σ Ac is associated to the dielectric relaxation caused by bound n. As reported in the literature [59,62,63], in the Non overlapping Small
charge carriers. Polaron tunnelling (NSPT), the exponent n increases with increasing
temperature. In the correlated Barrier Hopping (CBH), the exponent n
3.8.1. Ac-conductivity decreases as temperature increases. In the Overlapping-large Polaron
Fig. 9a displays the frequency dependence of Ac-conductivity for the Tunneling (OLPT), the exponent n decreases to a minimum value with
NBT sample over a wide temperature range 400–700 K. It is clearly seen increasing temperature and then begins to increase with rising tem­
that the Ac-conductivity curve may be divided into two regions; The perature. Finally, In the Quantum Mechanical Tunneling model (QMT),
first, corresponding to the Dc-conductivity, in low frequency range the exponent n is almost equal to 0.8 and increases with increasing
where the electric field cannot affect the jump conduction mechanism temperature or is independent of temperature.
and can be attributed to the electrode or surface effect. The second, Fig. 9c presents the variations of power exponents n1 and n2 with
corresponding to the alternating conductivity. It begins to increase temperature corresponding to the change in the conduction mechanism.
gradually at hopping frequency and can be attributed to the contribu­ As noticed the exponent ‘n1′ increased with temperature then decreased,
tions of grains and grains boundaries [58]. The rise in conductivity with suggesting that the small polaron hopping (SPH) and the correlated
frequency can be explained by the hopping of charge carriers between barrier hopping (CBH) model exist in this temperature range. In the case
localized sites [59]. The temperature range 400–700 K can be divided of ‘n2′ , it increased with temperature corresponding to the SPH model.

6
N. Rhimi et al. Inorganic Chemistry Communications 146 (2022) 110119

Fig. 8. (a) Dielectric constant vs frequency plot at different temperatures. (b)


Tan loss as a function of frequency for NBT ceramic.

Hence, according to these variants, SPH and CBH can explain the con­
duction mechanism.

3.8.2. Dc-conductivity
The data of σDC are fitted using Arrhenius equation [64]:

(14)
Ea
σ DC × T = eKB T
Where T is the absolute temperature, KB is Boltzmann’s constant and Fig. 9. (a) Frequency dependence of Ac-conductivity over a wide temperature
Ea is the activation energy. In this model, the activation energy and the range 400–700 K. (b) An example of fitting the experimental data for T = 600 K.
hopping range are independent of temperature. Fig. 10a displays the (c) The variations of power exponents n1 and n2 with temperature for
plot log(σDC × T) versus (1000/T) for the NBT ceramic. As observed, σDc NBT ceramic.
increased with temperature, showing that the bulk conduction is ther­
mally activated in NBT. The slope changed when the structural transi­ energy, Ea was calculated using the following expression [67]:
tion from the rhombohedral to tetragonal phase occured. Hence distinct
d(log(σ Dc ))
values of Ea . Generally the different values of Ea in different temperature Ea = − (15)
d(KB1 T )
regions in NBT are due to the difference in the conductivity mechanism
and domain configuration [65]. The value of Ea (0.255 eV) obtained in The inset of Fig. 10b shows the dependence of Ea on temperature.
the rhombohedral phase can be associated to the carrier transport pro­ Contrarily to the Arrhenius model, the activation energy changes with
cess associated with the jumping between localized states Ti4+ and Ti3+. temperature. Consequently, Mott assumed that the conduction is ther­
While the value of Ea (0.239 eV) obtained in the rhombohedral phase, mally activated by polaron tunnelling between localized states [68,69].
distorted by the tetragonal phase, can be due to the small polaron Mott’s Variable Range Hopping (VRH) equation is shown by the
created by the electron and/or hole–phonon interaction. In the tetrag­ following expression:
onal phase, the value of Ea (0.139 eV) can be due to the oxygen vacancies
[66,67].
1
(16)
T
( T0 )4
σ Dc = σ 0 e−
To prevent the impact of the conduction model on the activation

7
N. Rhimi et al. Inorganic Chemistry Communications 146 (2022) 110119

value to be 6 Å [71]. Fig. 10b displays the log(σDc ) vs T − 4 plot for the NBT
1

ceramic. The value of N(EF), calculated from Mott’s characteristic


temperature, was utilized to calculate the hopping distance DH and the
activation energy Ea by the following equations [43,72]:
[ ]14

DH = (18)
8π KB TN(EF )
And.
3
Ea = (19)
4πD3H N(EF )

The calculated Mott’s parameters for NBT are detailed in Table 1.


Fig. 10c displays the variation of activation energy and distance DH with
temperature. For the three regions, the hopping distance decreased from
7.86 to 3.75 nm as temperature increased from 400 to 700 K where the
minimum value of DH was at around 700 K, suggesting that this tem­
perature presents the maximum temperature corresponding to the
shortest hopping distance permitted for the occurrence of the VRH
process [73].

4. Conclusion

Dense NBT ceramic was developed successfully by solid-state


method. The structural, optical and electrical properties were investi­
gated. The Rietveld refinement showed rhombohedral structure with
R3c space group with the presence of small amount of Na0.5Bi4.5Ti4O15
as impurity. From the SEM analysis, the average calculated grains size
was 29.17 µm. The presence of different chemical bonds in the sample
was confirmed by FT-IR spectroscopy. Based on diffuse reflectance
spectroscopy, we determined the optical band gap energy, which was
2.61 eV. Complex impedance and modulus study revealed the presence
of both grain boundary and grain contributions at higher frequency
region and electrode/interfacial effects at lower frequency region in
NBT. An equivalent circuit was suggested for the electrical response of
the material and the relaxation effect in NBT was found to be non-Debye
type. Dielectric properties analysis exhibited a dielectric permittivity
and loss ε = 3871 and Tanδ = 0.995 at 100 kHz at TC = 580 K,
respectively. The dispersion of Ac-conductivity showed Jonscher’s
power law feature in the temperature range (400–460 K) and Jonscher’s
double power law in the temperature range (480–700 K). According to
the variation of n1 and n2 exponent, SPH and CBH could explain the
conduction mechanism. The temperature dependence of Dc-
conductivity of NBT followed Mott’s Variable Range Hopping (VRH)
law.

CRediT authorship contribution statement

Najah Rhimi: Conceptualization, Writing – original draft. N.


Dhahri: Data curation, Resources. M. Khelifi: Writing – review &
editing. E.K. Hlil: Data curation, Resources. J. Dhahri: Writing – review
& editing.
1
Fig. 10. (a) log(σ DC × T) versus (1000/T). (b)log(σ Dc ) versus T − 4 . (c) The Declaration of Competing Interest
variation of activation energy and distance DH with temperature for
NBT ceramic.
The authors declare that they have no known competing financial
interests or personal relationships that could have appeared to influence
Where σ 0 is a constant and T0 is Mott’s characteristic temperature, the work reported in this paper.
given by the following expression [42]:
24 Data availability
T0 = (17)
πK B N(EF )β3
Data will be made available on request.
Where N(EF) is the density of states at the Fermi level and β is the
decay factor of the localized polaron wave function. It can be considered
as the distance between the two nearest B-site atoms for perovskite
materials [70]. For the NBT perovskite, Pattipaka et al. reported this

8
N. Rhimi et al. Inorganic Chemistry Communications 146 (2022) 110119

References [27] W. Ge, C.P. Devreugd, D. Phelan, Q. Zhang, M. Ahart, J. Li, H. Luo, L.A. Boatner,
D. Viehland, P.M. Gehring, Lead-free and lead-based ABO3 perovskite relaxors with
mixed-valence A-site and B-site disorder: comparative neutron scattering structural
[1] X. Mao, Z. Li, M. Li, X. Xu, C. Yan, Z. Zhu, A. Du, Computational design and
study of (Na1/2Bi1/2)TiO3 and Pb(Mg1/3Nb2/3)O3, Phys. Rev. B. 88 (2013), 174115.
experimental validation of the optimal bimetal doped SrCoO3− δ perovskite as
[28] S.B. Vakhrushev, V.A. Isupov, B.E. Kvyatkovsky, N.M. Okuneva, I.P. Pronin, G.
solid oxide fuel cell cathode, J. Am. Chem. Soc. 143 (2021) 9507–9514.
A. Smolensky, P.P. Syrnikov, Ferroelectrics 63 (1985) 153–160.
[2] K. Ramam, S. Surabhi, S.C. Gurumurthy, M.P. Shilpa, K. Bindu, Ravikirana,
[29] B.K. Barick, R.N.P. Choudhary, D.K. Pradhan, Phase transition and electrical
S. Mundinamani, Dielectric and piezoelectric studies of dysprosium-doped BZT-
properties of lanthanum-modified sodium bismuth titanate, Mater. Chem. Phys.
BCNT perovskite ceramic system for sensors and actuator applications, J. Mater.
132 (2-3) (2012) 1007–1014.
Sci: Mater. Electron. 32 (13) (2021) 18002–18011.
[30] E. Aksel, J.S. Forrester, B. Kowalski, J.L. Jones, P.A. Thomas, Phase transition
[3] M.R. Kannan, A. Logeswari, M.W. Carry, T. Vijayakumar, Synthesis and
sequence in sodium bismuth titanate observed using high-resolution x-ray
investigation of (1–x)K0.5Na0.5NbO3–(x)CaSnO3 lead free perovskite ceramics of
diffraction, Appl. Phys. Lett. 99 (2011), 222901.
high dielectric and piezoelectric properties for transducer applications, J. Mater.
[31] Y. Guo, H. Fan, C. Long, J. Shi, L. Yang, S. Lei, Electromechanical and electrical
Sci: Mater. Electron. 33 (2022) 9224–9234.
properties of Bi0.5Na0.5Ti1-xMnxO3-δ ceramics with high remnant polarization,
[4] O. Amhoud, N. Zaim, M. Kerouad, A. Zaim, Effect of Cd2+ element substitution at
J. Alloys Compd. 610 (2014) 189–195.
the Nd3+ site on the magnetic, the magnetocaloric and hysteresis properties of Nd1-
[32] B. Parija, T. Badapanda, S.K. Rout, L.S. Cavalcante, S. Panigrahi, E. Longo, N.
x CdxMnO3 perovskite: Monte Carlo study, Mater. Today. Proc. 45 (2021)
C. Batista, T.P. Sinha, Morphotropic phase boundary and electrical properties of
7531–7537.
1–x [Bi0.5Na0.5]TiO3-x Ba[Zr0.25Ti0.75]O3 lead-free piezoelectric ceramics, Ceram.
[5] Y.H. Ko, M. Jalalah, S.J. Lee, J.G. Park, Super ultra-high resolution liquid-crystal-
Int. 39 (5) (2013) 4877–4886.
display using perovskite quantum-dot functional color-filters, Sci. Rep. 8 (2018)
[33] B.K. Barick, R.N.P. Choudhary, D.K. Pradhan, Dielectric and impedance
1–7.
spectroscopy of zirconium modified (Na0.5Bi0.5)TiO3 ceramics, Ceram. Int. 39 (5)
[6] H. Schlenz, S. Baumann, W.A. Meulenberg, O. Guillon, The development of new
(2013) 5695–5704.
perovskite-type oxygen transport membranes using machine learning, Crystals 12
[34] M. Otoničar, S.D. Škapin, M. Spreitzer, D. Suvorov, Compositional range and
(12) (2022) 1–17.
electrical properties of the morphotropic phase boundary in the Na0.5Bi0.5TiO3-
[7] B.R. Sutherland, S. Hoogland, M.M. Adachi, C.T.O. Wong, E.H. Sargent, Conformal
K0.5Bi0.5TiO3 system, J. Eur. Ceram. Soc. 30 (4) (2010) 971–979.
organohalide perovskites enable lasing on spherical resonators, ACS Nano 8 (2014)
[35] Y. Pu, L. Zhang, Y. Cui, M. Chen, High energy storage density and optical
10947–10952.
transparency of microwave sintered homogeneous (Na0.5Bi0.5)(1–x)BaxTi(1–y)SnyO3
[8] N. Humera, F. Arshad, A. Raza, M.A. Raza, S. Atiq, Z.N. Kayani, S. Naseem, S. Riaz,
ceramics, ACS Sustain. Chem. Eng. 6 (5) (2018) 6102–6109.
Dielectric and ferroelectric properties of X8R perovskite barium titanate for
[36] H. Xie, L.i. Jin, D. Shen, X. Wang, G. Shen, Morphotropic phase boundary,
application in multilayered ceramics capacitors, J. Mater. Sci: Mater. Electron. 33
segregation effect and crystal growth in the NBT-KBT system, J. Cryst. Growth. 311
(10) (2022) 7405–7422.
(14) (2009) 3626–3630.
[9] J.H. Heo, D.H. Shin, S.H. Moon, M.H. Lee, D.H. Kim, S.H. Oh, W. Jo, S.H. Im,
[37] G.O. Jones, P.A. Thomas, Investigation of the structure and phase transitions in the
Memory effect behavior with respect to the crystal grain size in the organic-
novel A-site substituted distorted perovskite compound Na0.5Bi0.5TiO3, Acta
inorganic hybrid perovskite nonvolatile resistive random access memory, Sci. Rep.
Crystallogr. Sect. B 58 (2) (2002) 168–178.
7 (2017) 1–8.
[38] C.J. Bartel, C. Sutton, B.R. Goldsmith, R. Ouyang, C.B. Musgrave, L.M. Ghiringhelli,
[10] D. Yu, A. Kumar, T.A. Nguyen, M.T. Nazir, G. Yasin, High-voltage and ultrastable
M. Scheffler, New tolerance factor to predict the stability of perovskite oxides and
aqueous zinc− iodine battery enabled by N-doped carbon materials: revealing the
halides, Sci. Adv. 5 (2019) eaav0693.
contributions of nitrogen configurations, ACS Sustainable Chem. Eng. 8 (2020)
[39] H. Kacem, A. Dhahri, M.A. Gdaiem, Z. Sassi, L. Seveyrat, L. Lebrun, V. Perrin,
13769–13776.
J. Dhahri, Electrocaloric properties of lead-free ferroelectric ceramic near room
[11] G. Yasin, M. Arif, T. Mehtab, X. Lu, D. Yu, N. Muhammad, M.T. Nazir, H. Song,
temperature, Appl. Phys. A Mater. Sci. Process. 127 (2021) 1–10.
Understanding and suppression strategies toward stable Li metal anode for safe
[40] S. Devesa, A.P. Rooney, M.P. Graça, D. Cooper, L.C. Costa, Williamson-hall analysis
lithium batteries, Energy Storage Mater. 25 (2020) 644–678.
in estimation of crystallite size and lattice strain in Bi1.34Fe0.66Nb1.34O6.35 prepared
[12] H. Wang, L. Sheng, G. Yasin, L. Wang, H. Xu, X. He, Reviewing the current status
by the sol-gel method, Mater. Sci. Eng. B. 263 (2021) 114830.
and development of polymer electrolytes for solid-state lithium batteries, Energy
[41] C.S. Devi, G.S. Kumar, G. Prasad, Control of ferroelectric phase transition in nano
Storage Mater. 33 (2020) 188–215.
particulate NBT – BT based ceramics, Mater. Sci. Eng. B. 178 (5) (2013) 283–292.
[13] G. Yasin, M. Arif, J. Ma, S. Ibraheem, D. Yu, L. Zhang, D. Liu, L. Dai, Inorg. Chem.
[42] R. Roy, A. Dutta, Structural, optical and enhanced electrical properties of
Front. 9 (2022) 1058–1069.
vanadium alloyed sodium bismuth titanate solid solution synthesized by a
[14] G. Yasin, S. Ibrahim, S. Ibraheem, S. Ali, R. Iqbal, A. Kumar, M. Tabish, Y. Slimani,
chemical-mechanical hybrid method, J. Alloys Compd. 843 (2020), 155999.
T.A. Nguyen, H. Xu, W. Zhao, J. Mater. Chem. A 9 (2021) 18222–18230.
[43] Y. Huang, L. Luo, J. Wang, Q. Zuo, Y. Yao, W. Li, The down-conversion and up-
[15] G. Yasin, M.A. Khan, W.Q. Khan, T. Mehtab, R.M. Korai, X. Lu, M.T. Nazir, M.
conversion photoluminescence properties of Na0.5Bi0.5TiO3: Yb3+/Pr3+ ceramics,
N. Zahid, Results Phys. 14 (2019), 102404.
Appl. Phys. 044101 (2015).
[16] T. Mehtab, G. Yasin, M. Arif, M. Shakeel, R.M. Korai, M. Nadeem, N. Muhammad,
[44] M. Zeng, S.W. Or, H.L.W. Chan, First-principles study on the electronic and optical
X. Lu, Understanding and suppression strategies toward stable Li metal, Energy
properties of Na0.5Bi0.5TiO3 lead-free piezoelectric crystal, J. Appl. Phys. 107
Storage Mater. 21 (2019) 632–646.
(2010), 043513.
[17] O.M. Hemeda, M.E.A. Eid, T. Sharshar, H.M. Ellabany, A.M.A. Henaish, Synthesis
[45] Z. Raddaoui, R. Lahouli, S.E.L. Kossi, J. Dhahri, K. Khirouni, K. Taibi, Structural
of nanometer-sized PbZrxTi1-xO3 for gamma-ray attenuation, J. Phys. Chem. Solids
and thermoelectric properties of Ba0.97Nd0.0267Ti0.95W0.05O3 ceramic, J. Alloys
148 (2021), 109688.
Compd. 765 (2018) 428–436.
[18] X.J. Wang, J.H. Huang, J. Wang, Experimental research on the response
[46] J. Hu, H. Qin, Giant magnetoimpedance effect in La0.7Ca0.3MnO3 under low
characteristics of PLZT ceramics, Smart Mater. Struct. 24 (7) (2015) 075017.
magnetic field, J. Magn. Magn. Mater. 231 (2001) 7–9.
[19] Q. Guo, L. Hou, F. Li, F. Xia, P. Wang, H. Hao, H. Sun, H. Liu, S. Zhang,
[47] F. Gaâbel, M. Khlifi, N. Hamdaoui, L. Beji, K. Taibi, J. Dhahri, Microstructural,
Investigation of dielectric and piezoelectric properties in aliovalent Eu3+ -modified
structural and dielectric analysis of Ni-doped CaCu3Ti4O12 ceramic with low
Pb(Mg1/3Nb2/3)O3-PbTiO3 ceramics, J. Am. Ceram. Soc. 102 (12) (2019)
dielectric loss, J. Mater. Sci. Mater. Electron. 30 (16) (2019) 14823–14833.
7428–7435.
[48] K.S. Rao, B. Tilak, K.C. Varada Rajulu, A. Swathi, H. Workineh, A diffuse phase
[20] J. Suchanicz, M. Nowakowska-Malczyk, A. Kania, A. Budziak, K. Kluczewska-
transition study on Ba2+ substituted (Na0.5Bi0.5)TiO3 ferroelectric ceramic,
Chmielarz, P. Czaja, D. Sitko, M. Sokolowski, A. Niewiadomski, T.V. Kruzina,
J. Alloys Compd. 509 (25) (2011) 7121–7129.
Effects of electric field poling on structural, thermal, vibrational, dielectric and
[49] I. Ahmad, M.J. Akhtar, M.M. Hasan, Impedance spectroscopic investigation of
ferroelectric properties of Na0.5Bi0.5TiO3 single crystals, J. Alloys Compd. 854
electro active regions, conduction mechanism and origin of colossal dielectric
(2021), 157227.
constant in Nd1-xSrxFeO3 (0.1 ≤ x ≤ 0.5), Mater. Res. Bull. 60 (2014) 474–484.
[21] Y. An, C. He, C. Deng, Z. Chen, H. Chen, T. Wu, Y. Lu, X. Gu, J. Wang, Y. Liu, Z. Li,
[50] V. Thakur, A. Singh, A.M. Awasthi, L. Singh, Temperature dependent electrical
Poling effect on optical and dielectric properties of Pr3+-doped Na0.5Bi0.5TiO3
transport characteristics of BaTiO3 modified lithium borate glasses, AIP Adv. 5
ferroelectric single crystal, Ceram. Int. 46 (4) (2020) 4664–4669.
(2015), 087110.
[22] A. Mishra, A. De, G. Abebe, G. Jafo, G.D. Adhikary, A. De, Effect of A-site off-
[51] R. Pattanayak, S. Panigrahi, T. Dash, R. Muduli, D. Behera, Electric transport
stoichiometry on the microstructural, structural, and electromechanical properties
properties study of bulk BaFe12O19 by complex impedance spectroscopy, Phys. B
of lead-free, J. Mater. Sci. Mater. Electron. 32 (2021) 12578–12593.
Phys. Condens. Matter. 474 (2015) 57–63.
[23] M. Benyoussef, M. Zannen, J. Belhadi, B. Manoun, Z. Kutnjak, D. Vengust,
[52] H.S. Mohanty, A. Kumar, B. Sahoo, P.K. Kurliya, D.K. Pradhan, Impedance
M. Spreitzer, M. El Marssi, A. Lahmar, Structural, dielectric, and ferroelectric
spectroscopic study on microwave sintered (1–x)Na0.5Bi0.5TiO3–x BaTiO3
properties of Na0.5(Bi1-xNdx)0.5TiO3 ceramics for energy storage and electrocaloric
ceramics, J. Mater. Sci.: Mater. Electron. 29 (8) (2018) 6966–6977.
applications, Ceram. Int. 47 (18) (2021) 26539–26551.
[53] A. Zaafouri, M. Megdiche, M. Gargouri, Studies of electric, dielectric, and
[24] R. Roukos, S.A. Dargham, J. Romanos, D. Chaumont, Detection of morphotropic
conduction mechanism by OLPT model of Li4P2O7, Ionics 21 (7) (2015)
phase boundary in A-site/Ca-substituted Na0.5Bi0.5TiO3 complex oxides
1867–1879.
ferroelectric system, J. Alloys Compd. 840 (2020), 155509.
[54] A. Kumari, S. Sanghi, A. Agarwal, O. Singh, Investigation of crystal structure,
[25] M. Zannen, A. Lahmar, M. Dietze, H. Khemakhem, A. Kabadou, M. Es-Souni,
dielectric properties, impedance spectroscopy and magnetic properties of (1–x)
Structural, optical, and electrical properties of Nd-doped Na0.5Bi0.5TiO3, Mater.
BaTiO3–(x)Ba0.9Ca0.1Fe12O19 multiferroic composites, Ceram. Int. 47 (16) (2021)
Chem. Phys. 134 (2-3) (2012) 829–833.
23088–23100.
[26] D. Rout, K.S. Moon, J. Park, S.J.L. Kang, High-temperature X-ray diffraction and
[55] C.G. Koops, On the dispersion of resistivity and dielectric constant of some
Raman scattering studies of Ba-doped (Na0.5Bi0.5)TiO3 Pb-free piezoceramics, Curr.
semiconductors at audio frequencies, Phys. Rev. 83 (1) (1951) 121–124.
Appl. Phys. 13 (2013) 1988–1994.

9
N. Rhimi et al. Inorganic Chemistry Communications 146 (2022) 110119

[56] Z. Xu, H. Qiang, Y. Chen, Z. Chen, Microstructure and enhanced dielectric [65] B.K. Barick, K.K. Mishra, A.K. Arora, R.N.P. Choudhary, D.K. Pradhan, Impedance
properties of yttrium and zirconium co-doped CaCu3Ti4O12 ceramics, Mater. Chem. and Raman spectroscopic studies of (Na0.5Bi0.5)TiO3, Phys. D. Appl. Phys. 44
Phys. 191 (2017) (2017) 1–5. (2011), 355402.
[57] N. Rezlescu, E. Rezlescu, Dielectric properties of copper containing ferrites, Phys. [66] O. Raymond, R. Font, N. Suárez-Almodovar, J. Portelles, J.M. Siqueiros,
Status Solidi (a) 23 (2) (1974) 575–582. Frequency-temperature response of ferroelectromagnetic Pb(Fe1/2Nb1/2)O3
[58] S.R. Elliott, A.c. conduction in amorphous chalcogenide and pnictide ceramics obtained by different precursors. PartI. Structural and thermo-electrical
semiconductors, Adv. Phys. 36 (2) (1987) 135–217. characterization, J. Appl. Phys. 97 (2005), 084107.
[59] M. Amara, J. Hellara, F. Bourguiba, J. Dhahri, A. Bchetnia, K. Taibi, E.K. Hlil, [67] M. Rudra, S. Halder, S. Saha, A. Dutta, T.P. Sinha, Temperature dependent
Structural, morphological, and dielectric properties of lead-free BaFe1/2Nb1/2O3 conductivity mechanisms observed in Pr2NiTiO6, Mater. Chem. Phys. 230 (2019)
based ceramics: toward a deeper understanding of the dielectric mechanisms, 277–286.
J. Mater. Sci. Mater. Electron. 33 (7) (2022) 3485–3500. [68] A. Sinha, A. Dutta, Structural interpretation of ion transport and small polaron
[60] A.K. Jonscher, The ‘universal’ dielectric response, Nature 267 (5613) (1977) hopping conduction in Gd substituted nickel nanoferrites, Phys. Status Solidi Appl.
673–679. Mater. Sci. 215 (2018) 1–8.
[61] H.S. Mohanty, A. Kumar, B. Sahoo, P.K. Kurliya, D.K. Pradhan, Impedance [69] N.F. Mott, Metal-insulator transitions, Pure Appl. Chem. 52 (1979) 65–72.
spectroscopic study on microwave sintered (1–x)Na0.5Bi0.5TiO3–x BaTiO3 ceramics, [70] H. Han, C. Davis, J.C. Nino, Variable range hopping conduction in BaTiO3 ceramics
J. Mater. Sci. Mater. Electron. 29 (2018) 6966–6977. exhibiting colossal permittivity, J. Phys. Chem. C 118 (17) (2014) 9137–9142.
[62] Y. Ben Taher, A. Oueslati, N.K. Maaloul, K. Khirouni, M. Gargouri, Conductivity [71] S. Pattipaka, A.R. James, P. Dobbidi, Dielectric, piezoelectric and variable range
study and correlated barrier hopping (CBH) conduction mechanism in diphosphate hopping conductivity studies of Bi0.5(Na, K)0.5TiO3 ceramics, J. Electron. Mater. 47
compound, Appl. Phys. A 120 (4) (2015) 1537–1543. (2018) 3876–3890.
[63] A. Ghosh, Frequency-dependent conductivity in bismuth-vanadate glassy [72] A. Mukherjee, S. Basu, P.K. Manna, S.M. Yusuf, M. Pal, Giant magnetodielectric
semiconductors, Phys. Rev. B 41 (1990) 1479. and enhanced multiferroic properties of Sm doped bismuth ferrite nanoparticles,
[64] C.h. Rayssi, S. El.Kossi, J. Dhahri, K. Khirouni, Colossal dielectric constant and non- J. Mater. Chem. C 3 (2015) 10715–10722.
debye type relaxor in Ca0.85Er0.1Ti1-xCo4x/3O3 (x=0.15 and 0.2) ceramic, J. Alloys [73] H. Böttger, V.V. Bryksin, Hopping conductivity in ordered and disordered systems
Compd. 759 (2018) 93–99. (III), Phys. Stat. Sol. (b) 113 (1) (1982) 9–49.

10

You might also like