1. Introduction
Titanates with the chemical expression M
2O·
nTiO
2 (where ‘M’ denotes elements such as Li, Na, K, etc., and
n = 2–8) have attracted researchers’ interest over the past decade because of their outstanding properties. Depending on the value of ‘
n’, these titanates can have a laminated or tunnel crystal structure comprising octahedral TiO
6, which shares edges with interposing cations. K
2Ti
2O
5 and K
2Ti
4O
9 structures are made up of layered foils of TiO
5 or TiO
6 octahedra connected to potassium atoms [
1]. Especially, potassium titanates, such as K
2Ti
6O
13, K
2Ti
2O
5, and K
2Ti
4O
9, are known for their high thermal durability, chemical resistivity, dispersibility, insubstantial metal ion evacuation, and ion exchangers, covering several applications [
2]. The photocatalytic activity of long K
2Ti
6O
13 whiskers synthesized via the traditional sol-gel process was examined for H
2 evolution and CO
2 photoreduction [
3]. The photocatalytic efficiency of carbon quantum dot-modified K
2Ti
6O
13 composite was examined via amoxicillin degradation under visible light irradiation. [
4]. K
2Ti
6O
13 is also used in metallic supports, such as copper, to improve wear and tear resistance [
5]; plastics to improve mechanical and dielectric properties [
6]; and car brake lining cushions as a replacement for carcinogenic asbestos [
7]. K
2Ti
6O
13 nanorods were produced and tested as a potassium-ion battery anode material for large-scale energy storage systems [
8]. Zhang et al. researched the photocatalytic and antibacterial performance of Cu-doped K
2Ti
6O
13 nanowires prepared using a combination of sol–gel and hydrothermal methods [
9]. In general, alkali titanates are employed as metal ion absorbents, and their cation transfer safeguards the environment from the deadly radiation of exceedingly radioactive fluid squanders [
10]. Nanomaterials of titanate also exhibited high performance in gas detection, photovoltaic cells, and cells of high-energy [
11]. Water bodies have been effectively cleared of harmful industrial pollutants, including methyl orange and methyl blue, by TiS
3 nanostructures. Using the chemical vapor transport method, J. Singh et al. discovered two-dimensional TiS
3 nanoribbons with outstanding mobility properties, and their photocatalytic activities have been investigated for possible environmental applications [
12].
Some alternative advantages of transition metal-doped K
2Ti
6O
13 over Pb-based perovskite ceramics have been reported previously [
13,
14,
15]. For instance, by doping divalent Cu
2+ with K
2Ti
6O
13 ceramics, generally, charge compensation progress may evolve oxygen vacancies (
, which are assertive charge carriers in ceramics), changing the ligand symmetry around the dopant and ultimately enhancing the dielectric properties of ceramics [
14]. Alternatively, in the perovskite lattice, it is believed that acceptor impurities can accelerate dielectric aging and lead to the formation of compensating positively charged
, which can impact device performance [
14]. PbO-based ceramics have excellent ferroelectric properties, but the volatile nature of deadly lead oxide during the synthesis process produces environmental pollution and induces inconsistency in composition. Because alkali titanates are Pb-free ceramics and have potential applications in industry, studies on alkali titanates have been encouraged. Various synthesis routes exist to prepare potassium hexa-titanate (K
2Ti
6O
13), such as the solid-state reaction method, the sol–gel method, the hydrothermal method, etc. Such a variety of preparations could also change alkaline titanate materials by doping them with aliovalent transition metal ions. Vikram et al. described the structural analysis, microtubular particles by SEM micrographs, and ferroelectric–paraelectric-type phase transition of Cu-doped K
2Ti
6O
13 ceramics [
14]. Vikram et al. focused on EPR spectroscopy in another study on Mn
x:K
2Ti
6O
13 lead-free ceramics, recognizing Mn
4+, Mn
3+, and Mn
2+ partial substitutions at Ti
4+ lattice sites; they identified defective accomplice dipoles exhibited in low-field EPR signals at different Mn doping concentrations [
16]. This study is aimed to find the positive temperature coefficient resistivity and temperature dependence dielectric of Co
x:K
2Ti
6O
13 ceramics for PTC thermistor application, which has not been previously reported according to our literature search. In this work, we report the structural properties and the temperature dependence of the dielectric constant and conductivity, along with an estimation of the ferroelectric phase transition temperature of Co
x:K
2Ti
6O
13 lead-free ceramics.
3. Results and Discussions
Figure 2 demonstrates the XRD patterns of K
2Ti
6O
13 and cobalt-doped K
2Ti
6O
13 (x = 0.05, 0.10, and 0.15 mole%, respectively) ceramics at room temperature. The XRD patterns were analyzed to identify the phase and lattice parameters of the samples. The peak positions of all samples were similar, and there were no secondary phases visible in the patterns, indicating that the K
2Ti
6O
13 matrix appeared in a single phase with Co-doping. The XRD pattern of different samples confirmed the monoclinic structure of Co:K
2Ti
6O
13 ceramic.
The room-temperature lattice constants of K
2Ti
6O
13 and Co: K
2Ti
6O
13 ceramics were calculated by [
15]:
where
dhkl is interplanar distance;
h,
k, and
l are the Bragg plane Miller indices; and
β is angular orientation. Various structural parameters are listed in
Table 1. The physical size of unit cells in a crystal lattice is referred to as the lattice constant. The lattice parameters of a doped material vary in comparison with the undoped material when a heteroatom with a lower or higher radius is doped in a lattice structure. The size variation of the unit cell is directly influenced by the size of the impure dopant. The variance is further enhanced by the doping percentage. Ti
4+ has an ionic radius of 0.68 Å, while Co
2+ has an ionic radius of 0.74 Å. In Co-doped K
2Ti
6O
13, Ti with a reduced radius is replaced by Co with a slightly larger radius, resulting in an increase in the lattice constant.
Table 1 shows the volume of unit cells for various samples. It was observed that increasing the doping percentage consistently increased the unit cell volume, as shown in
Figure 3.
The most apparent peak at θ = (200) may be used to calculate the crystallite size (D
hkl) of the identified crystalline phase using an amended Scherrer formula (200) [
17,
18,
19]:
Dhkl denotes the crystalline size, k denotes a dimension free-form factor with a value near unity (0.9), λ is the Cu Kα1 X-ray wavelength, β is the line broadening (FWHM), and θ is the Bragg angle. The grain size was measured along the (200) peak and found to increase with the amount of Co in the ceramics. Because of the difference in ionic radii between Co2+ and Ti4+ ions, it was also expected that Co: K2Ti6O13 would be formed with a large crystalline size due to more agglomeration.
The crystallinity (C) of any sample can be characterized as [
20]:
where A
cryst is the total area of all crystalline peaks in an X-ray diffraction pattern, and A
total is the total area of the diffraction pattern.
Figure 3 exhibits the crystallinity of Co
x:K
2Ti
6O
13 ceramics as a function of the cobalt substituent concentration.
Table 1 shows that the Co-doped K
2Ti
6O
13 ceramics were slightly more crystalline than K
2Ti
6O
13, and the crystallinity increased with increasing Co content. The slight elevation in crystallinity was due to the addition of a heavier element, such as Co, in place of Ti.
Figure 4 depicts the variation of the dielectric constant of Co
x:K
2Ti
6O
13 ceramics with temperature at frequencies of 100, 200, and 400 kHz. It can be seen that the dielectric constant rose with temperature up to a certain critical temperature and then started decreasing with a further increase in temperature. The Curie transition (T
c, ferroelectric–paraelectric type transition) temperature of K
2Ti
6O
13 ceramics is reported at 300 °C [
16]. Dielectric peaks of Co
0.05:K
2Ti
6O
13 were observed at temperatures of 335 °C, 348 °C, and 349 °C for 100 kHz, 200 kHz, and 400 kHz, correspondingly (
Figure 4a). The dielectric constant peak at T
c diminished, and the transition temperature moved to a higher value with a rising frequency, showing evidence of the relaxation behavior of the dielectric material [
21]. For Co
0.10:K
2Ti
6O
13 ceramic, two ferroelectric–paraelectric peaks appeared at 283 °C (bigger peak, T
C1) and 430 °C (smaller peak, T
C2). Moreover, Co
0.15:K
2Ti
6O
13 showed a low permittivity transition peak at a comparatively higher temperature of 387 °C, with zero frequency dispersion up to 260 °C, which was almost in the vicinity of the tangent loss peak (
Figure 4c). Changes in relaxation time and lattice disorder caused by dopant incorporation could be the cause of the peaks shifting to higher temperatures. It was observed that the dielectric constant of Co
x:K
2Ti
6O
13 ceramic increased with Co-doping for x = 0.05, 0.10, and then it diminished inevitably for x = 0.15 mole% compared with that of pure K
2Ti
6O
13 [
16]. Because Co
2+ has greater atomic polarizability than titanium, the dielectric constant of Co
x: K
2Ti
6O
13 rose with increased Co-doping (where x = 0.05, 0.10). However, at a concentration of x = 0.15, the porosity of the sample increased, which raised the resistivity and made polarization incredibly difficult and may have lowered the dielectric constant. The dielectric peaks observed to broaden and become disordered in the lattice due to Co
2+ ions may have been responsible for this broadening peak, prompting heterogeneity in the micro-composition. The incorporation of Co
2+ ions into the lattice induced oxygen vacancies (
), causing distortion of the surrounding volume and adjustment of the local fields. Additionally, a broadening of the peak happens due to non-uniform internal fields resulting from (Fe–
) defect-associated dipoles [
22]. The transition temperature and dielectric peak of Co
x:K
2Ti
6O
13 ceramics at frequencies of 100 kHz, 200 kHz, and 400 kHz are listed in
Table 2.
Figure 5 shows the variation of the dielectric loss of Co
x:K
2Ti
6O
13 ceramics with temperature at frequencies of 100, 200, and 400 kHz. Dielectric loss was found to rise with temperature up to a certain temperature, and after that, it decreased with an increasing frequency for all doped compositions. Tangent loss peaks of Co
0.05:K
2Ti
6O
13 were observed at temperatures of 335 °C, 340 °C, and 347 °C for 100 kHz, 200 kHz, and 400 kHz, respectively. Herein, the tangent loss peak was moved to a higher temperature, and loss was reduced with increasing frequency, exhibiting ferroelectric relaxation that could be assigned to a reduction in the relaxation time with temperature [
23]. For Co
0.10:K
2Ti
6O
13 ceramic, the single ferroelectric–paraelectric peak bifurcated itself into two peaks, which occurred at 283 °C and 430 °C, following a trend similar to that of dielectric permittivity (
Figure 5b). However, a single transition peak was subsequently obtained at a temperature of 384 °C for Co
0.15:K
2Ti
6O
13 ceramic. The typical increment in tanδ with temperature might have been because of leakage current and an increase in the number of ions participating in polarization [
24]. The high rate of tangent loss at lower frequencies and higher temperatures results from space charge polarization [
25]. Compared with K
2Ti
6O
13 [
17]
, a nearly uniform dielectric loss pattern was obtained in doped Co
x:K
2Ti
6O
13. The peaks of tangent loss in various doped samples are credited to the homogeneity in the composition because of doping. However, the oxygen vacancies that keep the crystal structure stable also affect mobility in the oxygen sublattice (CoO
6 and TiO
6), acting as fast ion conductors [
26]. The origin of the loss peaks is expected to be the relaxation of permanent dipoles and (Fe–
) deformity associated with the dipoles [
25]. The dielectric behavior of Co
x:K
2Ti
6O
13 ceramics near phase transition (Tc) can be explained by utilizing the perovskite material idea, while the phase transition can be ascribed to the instability of the temperature-dependent soft phonon mode (i.e., low-frequency mode), in which the frequency approaches 0 and the corresponding lattice displacements become unstable [
23]. The soft mode concept may be used to explain the above properties using the real part of the dielectric constant and tangent loss, which are similar to the Debye equations [
27]:
and in the range
.
where
(q) denotes relaxation time and
(q) denotes the phonon frequency [
28]. Although the over-damped soft phonon mode temperature and dielectric relaxation behavior ended up being practically indistinguishable, the permittivity of the ceramics exhibits an abrupt change at the transition temperature (T
C) when the soft phonon mode damping and relaxation impact were insignificant. Moreover, term {1 +
(q)
(q)} in the denominator of the above expression alters the sharp transition to a rounded hump. The maximum loss and relaxation effect could be attributed to the condition
(q) = 1/
(q). The origin of phase transitions indicating dielectric irregularity originated from the unit cell structure, which is temperature-sensitive, and transformed into a progressively steady structure at a particular transition temperature. When the lattice is not able to relax against lattice stresses, assemblies of TiO
6 are intensely contorted, and Ti
4+ ions lie at its center of mass in an intact oxygen octahedron, developing permanent dipoles. Generally, these dipoles rearrange themselves toward the applied field, resulting in a substantial polarization influence and, consequently, spontaneous polarization in all compositions. It can be understood from
Figure 5 that the variation of tangent loss with temperature shows an extremely large energy loss near transition (but below T
c). Intense molecular thermal vibrations above T
c destroy spontaneous polarization, bringing decay in permittivity. Thus, the ferroelectric–paraelectric-type phase transition was confirmed by the observed peak of the relative permittivity (ε
r) at T
c.
Figure 6 illustrates the variation of ac conductivity (
σac) with temperature in the range of 100–500 °C for the frequencies of 100, 200, and 400 kHz. All samples have slightly unusual patterns of ac thermal behavior. On the basis of this behavior, the trend may be divided into four regions for analysis. The activation energy could be determined in different regions using Arrhenius equations [
29]:
where
𝜎o denotes the pre-exponential factor,
Ea denotes activation energy,
k denotes the Boltzmann constant, and
T denotes absolute temperature.
Ea can be calculated by multiplying the Boltzmann constant (0.00008617 eV/K) by the slope of the line.
Region 1 (100–250 °C). The low activation energy region observed in
Figure 6 depends on the frequency and temperature of all doped samples, while it slightly depends on temperature.
σac has the behavior of the Jonscher power law (
σac = Aω
s, where s < 1) and is typically assigned to the hopping motion of electrons [
29]. The variation trend of
σac with temperature was similar for all samples, but slightly more temperature dependence was observed for the x = 0.10 sample. The small incline (slope) in this region showed low activation energy for electric transport.
Region 2 (250–350 °C). In this region of temperature,
σac strongly depends on temperature and revealed an activation energy value higher than the temperature region below 250 °C, showing the appearance of interchangeable intratunnel ionic conduction. Multiple hops of the hopping electrons are responsible for the temperature dependency of ac conductivity [
29] that varies linearly with an improvement in activation energy. Indeed, in this region, the highest activation energy was observed for Co
0.10:K
2Ti
6O
13 ceramics.
Region 3 (350–420 °C). Conductivity showed anomalous behavior in this temperature region, showing relaxation peaks at 415 °C (for x = 0.05), 384 °C (for x = 0.10), and 407 °C (for x = 0.15). These relaxation peaks come into existence because of the domination of strong dipole alignment with enhanced doping. The shifting of peaks for different doped samples was due to changes in relaxation time and disorder in the lattice. Conductivity in this area rose to the transition temperature because more conduction electrons are created when an acceptor is ionized. After Tc, conductivity decreased a lot, which may be because the value of mobility goes down as the temperature goes up.
Region 4 (420–500 °C). Conductivity depends strongly on frequency and temperature in this temperature region. The average ac conductivity value in this region again began to increase after a specific temperature above the transition point. The sharp rise in conductivity in this region may have been due to a large increase in intrinsic conductivity, which compensated for the loss in mobility. Grain morphology and grain porosity also have a significant impact on conductivity. The material close to the transition temperature shows a positive temperature coefficient of resistivity (PTCR). The grain boundary is closely related to PTCR behavior [
30]. The PTCR effect is caused by trapped electrons at the grain boundaries, as explained by the well-established Heyyang model. The PTCR is a consequence of the dependence of barrier heights on dielectric constants above the transition temperature (T > T
c). The exponential increase in dielectric constants above T
c prompts a reduction in barrier heights and an exponential increase in conductivity. Grain boundary barriers do not have a clear significant role below T
c, and the reduction in barrier height is credited to the remuneration of the grain boundary charge by spontaneous polarization [
31]. At the point where the dielectric reduces with temperature, however close to the Tc, the barrier height rises, promoting an increase in resistivity with temperature up to the potential barrier boundary, which is equivalent to the activation energy. Regarding semiconductors, after that temperature, the resistivity begins to diminish with a further rise in temperature. Here, conductivity exhibited a strong temperature dependence and displayed an activation energy value, which is a significant novelty in this study.
The complex electric modulus formalism, which is based on polarization analysis, can also be used to evaluate the electrical behavior of materials [
32]. The electric modulus may reveal important information about the relaxation mechanism. This method has the benefit of minimizing grain boundary conduction, electrode effects, and the undesirable effects of extrinsic relaxation [
33]. The electrical modulus (M*), which is based on the complex permittivity ε* = ε′ − iε″, can be written as follows:
The imaginary and real components of the electrical modulus are M′′ and M′, respectively. The temperature dependence of real part M′ of the electrical modulus at 100 kHz, 200 kHz, and 400 kHz is shown in
Figure 7. It was shown that M′ decreased with temperature up to a certain temperature, and then it increased again with temperature, resulting in peaks at different temperatures. The values of M′ at 400 kHz for all doped samples were higher in almost the entire temperature range. In the higher temperature range, we observed one peak of M′ for the x = 0.05- and 0.15-doped samples, and double peaks were observed for the x = 0.10-doped samples. As shown in the dielectric plot in
Figure 4, we also observed double peaks for the x = 0.10-doped sample. At very high temperatures, the M’ values were close to zero, indicating that electronic polarization was minimal [
34].
Figure 8 shows the variation of the imaginary portion of the electrical modulus (M′′) with temperature at 100 kHz, 200 kHz, and 400 kHz. It was observed that values of M′′ at lower temperatures were lower, and they initially increased with temperature and then decreased, following the exhibition of peaks at higher temperatures. The values of M′′ at frequencies of 100 kHz, 200 kHz, and 400 kHz were almost the same for the whole temperature range. M′′ started increasing with temperature at lower temperatures for samples x = 0.05 and x = 0.10, whereas it was almost constant up to 225 °C for x = 0.15.