Next Article in Journal
Surface Modification of Bi2Te3 Nanoplates Deposited with Tin, Palladium, and Tin/Palladium Using Electroless Deposition
Previous Article in Journal
On Melt Growth and Microstructure Characterization of Magnesium Bicrystals
Previous Article in Special Issue
Glassy Properties of the Lead-Free Isovalent Relaxor BaZr0.4Ti0.6O3
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Impedance Spectroscopy of Fe and La-Doped BaTiO3 Ceramics

by
Malgorzata Adamczyk-Habrajska
,
Beata Wodecka-Dus
*,
Tomasz Goryczka
and
Jolanta Makowska
Faculty of Science and Technology, Institute of Materials Engineering, University of Silesia, 75 Pułku Piechoty 1A, 41-500 Chorzow, Poland
*
Author to whom correspondence should be addressed.
Crystals 2024, 14(2), 131; https://doi.org/10.3390/cryst14020131
Submission received: 27 December 2023 / Revised: 23 January 2024 / Accepted: 24 January 2024 / Published: 27 January 2024
(This article belongs to the Special Issue Advanced Electronic Ceramics)

Abstract

:
A wide range of the interesting properties of electroceramics Ba0.996La0.004Ti0.999O3 (BLT4) undoubtedly deserves differentiation and optimization. For this purpose, the corresponding donor oxide dope Fe2O3 was introduced in excess quantities into the base ceramics. In this way, an innovative ceramic material with the general formula of Ba0.996La0.004Ti1−yFeyO3 (BLTF), for y = 0.001, 0.002, 0.003, 0.004, has been produced. The crystal structure of BLTF ceramics was investigated using X-ray diffraction. The diffraction peaks in XRD confirm the formation of the tetragonal perovskite phase. The electrical properties of BLTF ceramics have been tested using impedance spectroscopy, in the frequency range of 20 Hz–2 MHz and the temperature range of 20–580 °C. To gain absolute certainty on the consistency of the measured data, the obtained impedance spectra were analyzed using the Kramers–Kronig method. The usage of an equivalent circuit, proposed by the authors, allowed grain and grain boundary resistivity to be obtained. Based on the diagram of the natural logarithm of the mentioned resistivity versus the reciprocal absolute temperature, the activation energies of the conductivity processes have been determined. The values of activation energies indicated that the admixture of iron introduced into the BLT4 ceramics played a crucial role in the conductivity of grains and intergranular borders.

1. Introduction

In the family of ceramic intelligent materials, a special place is occupied by ferroelectrics with a perovskite structure type ABO3 [1,2,3]. One of the representatives of these materials is barium titanate, an important perovskite ferroelectric material, which has been known since 1940 [4,5,6,7]. This fact is undoubtedly related to the wide range of application possibilities and considering the principles of the so-called green technologies, environmentally friendly. There are many methods for obtaining BaTiO3 ceramics, e.g., the free sintering method, the sol-gel method, or spark plasma sintering, which has been gaining attention in materials science research [8]. Barium titanate has been the most widely studied classical ferroelectric due to its excellent dielectric, ferroelectric, optical, piezoelectric, and other useful properties [9,10,11]. BaTiO3 finds practical applications in classic capacitors, thermistors, and varistors [12,13,14,15], piezoelectric sensors, and piezoelectric transducers [16,17,18]. Multilayer ceramic capacitors (MLCCs) with a BaTiO3 ferroelectric as the core material have a low-temperature coefficient of capacitance over a wide temperature range, and are currently the commercial MLCC with the largest demand in the market [19,20]. Ceramic materials based on BaTiO3 are also used to build ferroelectric random access memories; the so-called FeRAM [21].
The influence of iron ions on the structure, microstructure, and physical properties of BaTiO3 ceramics has been widely discussed in recent years [22,23,24]. However, there are no reports in the world literature on the influence of iron admixture on the electrical properties of BLT. In the article, the authors draw attention to changes in electrical phenomena. The literature on the subject shows that barium titanate is a relatively easily modified compound, and one of the most effective additives is lanthanum [25]. The authors [26,27,28] indicate that doping barium titanate with lanthanum ions in an amount of 0.4 mol%. (BLT4) leads to obtaining a ceramic material with a particularly high electrical permittivity value and interesting electrical properties. However, the admixture of iron introduced into BaTiO3 causes ferromagnetic properties and the appearance of high leakage currents [29]. In light of current research, it seems reasonable to dope BT ceramics simultaneously with lanthanum and iron ions. When considering how to introduce both dopants into the crystal lattice, their ion radii should be taken into account. The ion radius of La3+ (r = 1.216 Å) is similar to Ba2+ (r = 1.47 Å) in size. The surplus charge of La3+ ions is compensated, among others, by creating barium vacancies [30]. This substitution method is confirmed by the increase in electrical permeability noted by both the authors of this paper [31] and other scientists [32,33]. Similarly, Fe3+ (0.61 Å) does not enter the barium sites due to its radius but replaces the Ti4+ ions (0.68 Å). The difference in the charge of the discussed ions is compensated for by creating oxygen vacancies [30]. Moreover, introducing La3+ ions into the Ba2+ positions reduces the number of covalent bonds, and substituting Fe3+ ions in place of Ti4+ ions leads to a reduction in the movement of oxygen octahedra. This may result in a high possibility of regulating the temperature of the ferroelectric phase transition and a significant increase in the dielectric constant of this compound at room temperature (Tr). The influence of iron ions on the structure, microstructure, and physical properties of BaTiO3 ceramics has been widely discussed in recent years [34,35,36,37]. Nakayama and others pointed out that the Fe-doped BaTiO3 ceramics are naturally ferromagnetic (FM) [38]. Another work investigated the simultaneous substitution of iron at the A and B sites in BaTiO3. It was found that when Fe was substituted at the B-site, it improved the coercive field, while at the A-site substitution, sample saturation magnetization increased [38]. However, there is very little information in the world literature on the effect of iron admixture on the electrical properties of BLT.
The authors of the present paper started the investigation on barium titanate modified simultaneously by La3+ and Fe3+ ions a couple of years ago. The previous paper widely described the influence of simultaneous modifications on electrical and dielectric properties caused by introducing the La3+ and Fe3+ ions to the crystal lattice [39]. In another paper, the authors described changes in the piezoelectric and elastic properties and increased barium titanate conductivity [40].
In the article, the authors draw attention to changes in electrical phenomena occurring in BLT4 ceramics under the influence of the iron admixture. Iron ions replace Ti sites in the perovskite structure and change the space charge concentration. The dopant modifies the process of accumulation and migration of charged defects and species to the grain boundary [37]. A comprehensive impedance analysis was performed to obtain information about the electrical properties of the tested material. This technique allows for the separation of the individual contribution of grains and grain boundaries in the total impedance spectrum. Also, it allows the examination of the influence of the iron admixture on the dynamics of both bound and free charges inside and in the electrode layers of the ceramic material [41].

2. Materials and Methods

Iron dopant was added to the base material Ba0.996La0.004Ti0.999O3 (BLT4). The modified ceramic materials with the general formula of Ba0.996La0.004Ti1−yFeyO3 (BLTF), for y = 0.1 mol.% (BLTF1), 0.2 mol.% (BLTF2), 0.3 mol.% (BLTF3), and 0.4 mol.% (BLTF4), were obtained using the conventional mixed oxides method. The proper amounts of substrates: BaCO3, La2O3, TiO2, and Fe2O3 (purity ≥ 99%) were weighed in a stoichiometric ratio considering the expected chemical reaction (1):
(1−x)BaCO3 + (1/2x)La2O3 + (1−y)TiO2 + (1/2y)Fe2O3→Ba1−xLaxTi1−yFeyO3 + (1−x)CO2
The stoichiometric formulas of discussed ceramics contain the real amount of ingredients:
Ba0.996La0.004Ti0.999Fe0.001O3⇒ BLTF1;
Ba0.996La0.004Ti0.998Fe0.002O3⇒ BLTF2;
Ba0.996La0.004Ti0.997Fe0.003O3⇒ BLTF3;
Ba0.996La0.004Ti0.996Fe0.004O3 ⇒ BLTF4.
The mixture of powders with the addition of ethyl alcohol C2H5OH was subjected to grinding in a high-energy ball mill (300 rpm) for 8 h, using zirconium yttrium balls as grinders. The synthesis was conducted at T = 950 °C in a corundum crucible in an air atmosphere for 6 h. The synthesized ceramic materials were then thoroughly ground, pressed and formed into disks with the diameter of d = 10 mm and the thickness of p = 1 mm, under the pressure of p = 30 MPa. BLTF ceramics were sintered at the temperature of TI = 1250 °C, TII = 1300 °C, and TIII = 1350 °C for 2 h. All samples were furnace-cooled to room temperature. Basic material parameters, namely real density obtained using the Archimedes method with distilled water and porosity, are summarized in Table 1. Previous studies have shown that incorporating a 4 mol.% lanthanum dopant significantly affects the density of BaTiO3 ceramics, causing an increase from the value of 5.39 g/cm3 (pure BaTiO3) to 5.63 g/cm3 (ceramic containing 0.4 mol% La). This problem was discussed in [42]. Introducing iron ions at concentrations equal to 0.1 and 0.2 mol% La slightly increased the density of the discussed material (Table 1); however, further increasing the dopant concentration led to a decrease in density to a value of 5.61 g/cm3. Similar behavior was observed in other materials based on iron-modified BaTiO3 ceramics [39]. The obtained real density is approximately equal to 95% of the theoretical density.
For electric measurements, the surfaces of the sintered pellets were ground and polished, followed by the application of a silver conducting paste on two opposite surfaces of the sample. The crystalline structure of the investigated ceramics was studied using the X-ray diffraction technique (XRD). The measurements were performed at room temperature with the use of the X’Pert Pro diffractometer equipped with CuKα1and2 radiation (40 kV, 30 mA). Diffraction patterns were measured in the step-scan mode in the angular range from 10° to 140° 2θ, with a measuring step of 0.05° and a counting time of 10 s per step. The measured diffraction patterns were fitted using the Rietveld method implemented in the LHPM computer program (version 4.2. Lucas Heights Research Laboratories ANSTO, Sydney, Australia) [43]. SEM studies of the BLTF ceramic samples, as well as analysis of their chemical composition, were performed by scanning electron microscopy SEM, JEOL JSM-7100 TTL, Tokyo, Japan equipped with an energy dispersive X-ray spectrometer (EDS)). The linear intercept length method was used to determine the grain size. The impedance studies were performed using a computerized measuring system, the integral part of which was an Agilent E4980A LCR meter. This system allowed the performance of complex impedance tests as a function of frequency (in the range of 20 Hz–2 MHz) and temperature (in the range from room temperature to 580 °C). The method using the K-K test V1.01 program confirmed the consistency of experimental data based on the Kramers–Kronig method [41,44]. The method is based on the Kramers–Kronig relationship, which is a very useful tool for data validation [45,46,47]. According to B.A. Boukamp [44], the dataset complied with the Kramers–Kronig transformation rules, which mean it must be:
-
casual, which means that the measured response is a consequence only of the applied signal,
-
a linear response, i.e., the system cannot generate an output frequency higher than the input frequency,
-
stable, which means that the system must be unchanged in the time, not continue to oscillate after the excitation is stopped,
-
finite for all frequencies, including ω→0 and ω→∞.
The Kramers–Kronig rules state that the imaginary part of impedance spectroscopy is completely determined by the form of the real part over the frequency range 0 ≤ ω ≤ ∞ and vice versa; the real part of impedance spectroscopy is determined by the imaginary part in the same frequency range. The sentence is described by Equations (2) and (3):
Z r e ω = R + 2 π 0 x Z i m x ω Z i m ω x 2 ω 2 d x
Z i m ω = 2 ω π Z r e x Z r e ω x 2 ω 2 d x
where Zre (ω)—is the real part of impedance obtained from the imaginary part,
Zim (ω)—is the imaginary part of impedance obtained from the real part.
Based on the values calculated from the presented equation as well as experimentally measured, the residuals could be defined in the following relationship (4) and (5):
r e , i = Z r e , i Z r e ω i , a k Z ω i , a k
i m , j = Z i m , J Z i m ω i , a k Z ω i , a k
For a good match between the data and the model, the residuals should be randomly distributed around the log(ω) axis, and their value should not exceed 5%.
A detailed analysis of the experimental data was performed using the Zview software (version 3.5, ScribnerAssociates Inc., Southern Pines, NC, USA).

3. Results

3.1. X-ray Diffraction

X-ray diffraction patterns of the investigated materials are presented in Figure 1. First, phase identification was performed using the ICDD (International Centre for Diffraction Data) PDF-4 database. For all samples, XRD diffraction patterns show the presence of a single perovskite phase without any secondary phase peaks. The layout of diffraction lines is characteristic of a tetragonal perovskite crystal structure with space group P4mm (ICDD card number 01-005-0626).
To explain the influence of Fe content on the crystallographic structure, the obtained results were subjected to refinement using the Rietveld method. Crystallographic data from ICDD PDF-4 file no. 01-005-0626 was used to build a theoretical model of the unit cell. By the assumptions of the chemical composition, the Ba positions were taken by La with an appropriate share modeled by the “site occupation” parameter. A similar approach was performed in the case of Ti, which Fe replaced. Lattice parameters calculated from refinement are collected in Table 2. The parameters of the unit cell practically do not change with the increase of iron dopant. The cell volume of the BLTF1 is VV = 64.3 × 1030 m3 and unit cell volume in BLTF4 is V = 64.2 × 10−30 m3. Obtained results show that the introduced dopants (La, Fe) led to the tetragonalization of the crystal lattice, as indicated by the parameter (δT). The increase in iron concentration causes a slight decrease in the δT parameter, which is caused by a decrease in the c0 parameter.

3.2. Microstructure and EDS Analysis

The study of Ba0.996La0.004Ti1−yFeyO3 ceramics morphology was carried out to determine the effect of Fe3+ admixture on the microstructure of the based ceramics. Figure 2 presents SEM images of BLTF ceramic fractures, for the various concentrations of added iron admixture.
The presented images indicated that BLTF ceramic materials are characterized by a densely packed, fine-grained microstructure with well-formed angular grains and a small number of pores. The results agree with the values of obtained density and porosity. The size of the grains increases with the increasing iron dopant concentration from d = 0.5 μm (for BLTF1) to d = 2 μm (for BLTF4). Therefore, the grain size of all discussed ceramics stays significantly lower than those observed in unmodified Ba0.996La0.004Ti0.999O3. The BLT4 ceramics are characterized by well-shaped large angular grains (d = 5 µm), with a tendency to spiral hexagonal growth [33]. The appearance of the microstructure indicates that, during the fractures, some cracks occurred through the grains (transcrystalline cracks). These kind of fractures were observed in the other materials [48,49,50]. This behavior indicates high mechanical strength in the grain boundaries at the expense of the interior of the grains [51].
The qualitative and quantitative chemical composition of all the discussed BLTF ceramics was tested with the energy-dispersive X-ray spectroscopy method. The obtained spectra are graphically presented in Figure 3, while for the pure compound in this work, see [14].
The spectra confirm the presence of La, Ti, Ba, Fe, and O in all the tested samples. No other element was detected. The graph shows the additional maximum, derived from the graphite layer sprayed on the sample. This maximum was not taken into consideration during the quantitative analysis. The chemical composition of the ceramic materials, determined on the basis of the results of the EDS analysis, remains in good agreement with the assumed stoichiometry. The percentage of individual components in the tested ceramics showed very slight deviations from the theoretically assumed quantities of oxides, and amounted to ±1 wt.%.

3.3. Impedance Analysis

Impedance analysis is a powerful technique that which provides information about the contribution of the microstructure components (such as the grain, grain boundary, and electrode interface) contribution to the electrical conductivity process. The technique was also used to determine the influence of the iron admixture on the electrical properties of the grains and grain boundaries of the BLT4 ceramics. The impedance spectra (SI) were determined at a wide range of temperatures for all of the discussed ceramics. Figure 4 presents frequency dependencies of real and imaginary impedance components at selected temperatures. The obtained characteristics have a smooth course, which indicates properly conducted measurements. In order to be absolutely certain about the consistency of the measured data, the obtained impedance spectra were analyzed using the Kramers–Kronig method. The sample residual spectra calculated for the temperature of 350 °C are shown in Figure 5. The test quality parameter was the coefficient χ2 [52].
The residual values of the results analyzed did not exceed 1%, whereas the distribution relative to the frequency axis was accidental, indicating the consistency of the measurement data. Further analysis of the impedance spectra is therefore most sensible.
The frequency characteristics of complex impedance for BLTF1 ceramics are significantly different from others presented. Namely, only one broadened maximum exists on the frequency dependencies of the imaginary part of impedance, and a single plateau is visible on the frequency dependencies of the real part of impedance. Moreover, the increase in the value of the real and imaginary part of the impedance of the discussed samples should be noted, in comparison to the values recorded for the basic ceramics BLT4. The shape of both discussed characteristics changes diametrically for the further considered concentrations of the admixture of iron. In the case of the BLT4 ceramics containing 0.2 mol.% of iron, the outline of the second maximum appears on the logZim(f) dependencies, in the range of high frequencies. The maximum becomes more visible for ceramics containing 0.3 and 0.4 mol.% Fe. Moreover, for ceramics with the admixture of iron exceeding 0.1 mol.%, a significant influence of temperature on the shape of the real and imaginary parts of the impedance spectra is visible.
To fully describe the electrically active areas of BLT ceramics doped with iron, the obtained results are presented in the form of the so-called Nyquist graphs (Figure 6), which significantly facilitates the selection of a suitable equivalent circuit.
The dependence of Zim(Zre) for ceramics containing 0.1 mol.% Fe obtained at different temperatures has the shape of deformed, symmetrical circles, the centers below the axis representing the real part of the impedance. The shape of the Nyquist chart, as well as the Zre(f) and Zim(f) dependencies pointing at the circuit consisting of single serially connected R CPE elements, is the best equivalent circuit to describe the electrical behavior of BLTF1 (Figure 7).
The CPE element has an impedance response that displays non-ideal frequency dependent properties and a constant phase over the entire range of frequencies. The impedance (Z) of CPE is defined by the CPE constant T (CPE-T) and CPE index p (CPE-P), which are parameters given by the following relation (6):
Z = 1 [ T j ω P ]
The CPE is identical to a capacitance component when the exponent p equals 1, and to a resistance component when p = 0 [53,54].
The complex impedance in the proposed electrical equivalent circuit is expressed by Equation (7):
Z = R 1 + T R ( j ω ) P
The result of the fitting procedure for several selected temperatures is shown in Figure 8. All fitted parameters for BLTF1 at selected temperatures are presented in Table 3, together with the errors of the fits for each parameter.
The obtained bulk resistance has been presented as lnR(1/T) dependence. The dependence is linear, which points the activation character of the conduction process (Figure 9). Based on the Arrhenius formula, the activation energies (Ea) of the process in question were determined. The value is equal to 1.03 eV.
In the case of the BLT4 ceramics modified with 0.2 mol.% of iron, the outline of the second semi-circle appears on Zim(Zre) dependencies in the high frequency range, which is connected with the electrical response of grains strongly suppressed by the electrical response of the grain boundaries With the further increase of iron concentration, the discussed semicircle, associated with the electrical response of grains, becomes more visible on the characteristics of Zim(Zre). Nevertheless it remains much smaller than the semi-circle representing the electrical response of the grain boundaries. The shapes Zim(Zre) dependencies pointing at the circuit consisting of two serially connected RC elements is the best equivalent circuit to describe the electrical behavior of BLT4 ceramics modified by iron ions. However, taking into consideration the value of the χ2 coefficient, the capacitance of both RC elements is replaced by CPE constant phase element (Figure 10).
The complex impedance in the proposed electrical equivalent circuit (Figure 10) is expressed by the Equation (8):
Z ω = Z G + Z G B
where (9), (10):
Z G = R G 1 + T G R G ( j ω ) P G
Z G B = R G B 1 + T G B R G B ( j ω ) P G B
The modification has significantly reduced the value of the χ2 coefficient, which means an improvement in the fitting quality.
The use of the above-described system allowed us to distinguish the contribution of grains and grain boundaries to electrical conductivity. The example result of the fitting procedure for BLTF4 ceramics at several selected temperatures is shown in Figure 11. All fitted parameters for BLTF2, BLTF3, and BLTF4 at selected temperatures are presented in Table 4, together with the errors of the fits for each individual parameter.
The obtained grain and grain boundary resistance in individual temperatures allowed us to determine the dependence of lnRG(1/T) and lnRGB(1/T) as shown in Figure 12.
All presented dependencies possess linear character, which has allowed us to determine the activation energy. The activation energy connected with the electrical conductivity of the grain boundary stays practically unchanged with the increase of iron concentration. The activation energy of grains is slightly less than the activation energy associated with the conductivity process in grain boundaries. Noteworthy is its sharp decline for the highest concentration of iron admixture. When interpreting the obtained impedance results, we must consider the direct impact of iron ions on the conductivity processes and the indirect impact, i.e., changes in the microstructure of the modified materials. The admixture of iron caused an increase in the grain size. In the undoped BaTiO3 material, increasing the grain size causes a significant decrease in both the grain resistance and the activation energy value [55]. In the case of the materials described in this work, the changes in activation energy are much more complicated. Namely, the admixture of iron causes an increase in the activation energy from the value of 0.87 eV (for the base Ba0.996La0.004Ti0.999O3 ceramics [33]) to the value of 1.03 eV (reported for ceramics containing 0.2 mol.% of iron). A further increase in iron content reduces the activation energy, but its value does not reach the initial value. However, the grain resistance shows an increasing tendency. The phenomena of electric charge transport are undoubtedly also influenced by the degree of porosity of the ceramic material. Based on the research presented by the authors of the work [56], significant changes in the values of electrical permittivity and loss tangent are visible only when the pore content exceeds 11%. These values are related to the commonly known relationships with complex impedance. Therefore, considering slight differences in the porosity of the materials presented in this paper, it can be assumed that the influence of porosity is negligible.
Considering the valence of iron ions and their ionic radius, it was assumed that they substitute in place of Ti4+ ions. The results of temperature tests of the tangent dependence of the loss angle, widely discussed in the paper [39], clearly confirmed the correctness of these assumptions. Namely, the tangent of the loss angle decreased with the increased concentration of iron ions. This change can be attributed to the decreasing number of unstable Ti4+ ions because Fe3+ occupying the position of Ti4+ ions reduces the possibility of changing the oxidation state of Ti4+ ions [57]. The reduction is responsible for the change in the concentration of space charge. The main reservoir of space charge is the grain boundaries. Therefore, changing the concentration of space charge should significantly change their resistance. Such changes occur in the case of discussed ceramic materials. The research results presented in this paper show that the resistance of grain boundaries increases rapidly. For example, grain boundary resistance increases at a temperature of 402 K from 26,620 Ω for unmodified ceramics [33] to 214,680 Ω for ceramics with 0.4 mol% Fe3+ ions. The grain resistance increase is another confirmation of the proposed explanation’s correctness. At a temperature of 402 K, the grain resistance changed from 2701 Ω for pure ceramics [33] to 9559 Ω for BLTF 4 ceramics.
Considering the previously published results of dielectric and electrical measurements [39] and the slight PTCR effect observed, the most promising material is BLT1 ceramics.
The research widely discussed in the present paper revealed that these ceramics are homogeneous from the point of view of conductivity—there is no distinction between grain conductivity and grain boundaries. This fact may be an additional advantage in the previously discussed applications.

4. Conclusions

The analysis of the microstructure made it possible to reveal a fine microstructure with well-formed, angular, and homogeneous crystalline grains containing dissolved donor admixtures, which confirmed that the sintering conditions used promoted the synthesis of the material. The tendency of the grains’ growth was observed with the increasing concentrations of the Fe3+ admixture. The elements of the microstructure, namely grains, and intergranular borders contribute to electrical conductivity. To determine the effect of iron on the electrical properties of BLT4 ceramics, impedance spectroscopy was used as a function of temperature, which confirmed the thesis of the complexity the mechanism of electric charge transport in the tested ceramic materials. On the basis of Nyquist’s graphs, the mutual overlap of two semi-circles connected with two microstructure components, namely grains and grain boundaries, was observed in the case of BLT4 ceramics containing 0.2, 0.3, and 0.4 mol.% of iron admixture. The dominant semi-circle is associated with the electrical response of the grain boundaries. The energy of activation of the conductivity of grains is smaller than the values of the activation energy associated with conduction in the grain boundary.

Author Contributions

Author contributions: Conceptualization, B.W.-D. and M.A.-H.; methodology, J.M.; software, M.A.-H. and T.G.; validation, B.W.-D.; formal analysis, B.W.-D., M.A.-H. and T.G.; investigation, B.W.-D.; resources, M.A.-H. and B.W.-D.; data curation, B.W.-D.; writing—original draft preparation, B.W.-D. and M.A.-H.; writing—review and editing, B.W.-D. and J.M.; visualization, B.W.-D.; supervision, M.A.-H.; project administration, M.A.-H.; funding acquisition, M.A.-H. All authors have read and agreed to the published version of the manuscript.

Funding

The present paper was financed in part by the Polish Ministry of Education and Science within statutory activity.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hańderek, J. Własności i Zastosowanie Wybranych Ferroelektryków; Prace Naukowe Uniwersytetu Śląskiego w Katowicach. Prace Fizyczne, nr 4; Wydawnictwo Uniwersytetu Śląskiego: Katowice, Poland, 1976. [Google Scholar]
  2. Smoleński, G.A.; Krajnik, N.N. Ferroelektryki i Antyferroelektryki; PWN: Warszawa, Poland, 1971. [Google Scholar]
  3. Ahmed, T.; Khan, S.A.; Kim, M.; Akram, F.; Park, H.W.; Hussain, A.; Qazi, I.; Lim, D.H.; Jeong, S.-J.; Song, T.K.; et al. Effective A-site modulation and crystal phase evolution for high ferro/piezoelectric performance in ABO3 compounds: Yttrium-doped BiFeO3-BaTiO3. J. Alloys Compd. 2023, 933, 167709. [Google Scholar] [CrossRef]
  4. Kinoshita, K.; Yamaji, A. Grain-size effects on dielectric properties in barium titanate ceramics. J. Appl. Phys. 1976, 47, 371–373. [Google Scholar] [CrossRef]
  5. Hiramatsu, T.; Tamura, T.; Wada, N.; Tamura, H.; Sakabe, Y. Effects of grain boundary on dielectric properties in fine-grained BaTiO3 ceramics. Mater. Sci. Eng. B 2005, 120, 55–58. [Google Scholar] [CrossRef]
  6. Uchino, K.; Sadanaga, E.; Oohashi, K.; Morohashi, T.; Yamamura, H. Particle/grain size dependence of ferroelectricity. Ceram. Trans. 1989, 8, 107–115. [Google Scholar]
  7. Arlt, G.; Hennings, D.; de With, G. Dielectric properties of fine-grained barium titanate ceramics. J. Appl. Phys. 1985, 58, 1619. [Google Scholar] [CrossRef]
  8. Bell, J.G.; Graule, T.; Stuer, M. Tuning of the microstructural and electrical properties of undoped BaTiO3 by spark plasma sintering. Open Ceramics 2022, 9, 100244. [Google Scholar] [CrossRef]
  9. Everhardt, A.S.; Denneulin, T.; Grünebohm, A.; Shao, Y.-T.; Ondrejkovic, P.; Zhou, S.; Domingo, N.; Catalan, G.; Hlinka, J.; Zuo, J.-M.; et al. Temperature-independent giant dielectric response in transitional BaTiO3 thin films. Appl. Phys. Rev. 2020, 7, 011402. [Google Scholar] [CrossRef]
  10. Pan, M.; Zhang, C.; Wang, J.; Chew, J.W.; Gao, G. Multifunctional Piezoelectric Heterostructure of BaTiO3@Graphene: Decomplexation of Cu-EDTA and Recovery of Cu. Environ. Sci. Technol. 2019, 53, 7923–8476. [Google Scholar] [CrossRef] [PubMed]
  11. Khan, S.; Humera, N.; Niaz, S.; Riaz, S.; Atiq, S.; Naseem, S. Simultaneous normal—Anomalous dielectric dispersion and room temperature ferroelectricity in CBD perovskite BaTiO3 thin films. J. Mater. Res. Technol. 2020, 9, 11439–11452. [Google Scholar] [CrossRef]
  12. Ianculescu; Mocanu, Z.V.; Curecheriu, L.P.; Mitoseriu, L.; Padurariu, L.; Trusca, R. Dielectric and tunability properties of La-doped BaTiO3 ceramics. J. Alloys Compd. 2011, 509, 10040–10049. [Google Scholar] [CrossRef]
  13. Petrović, M.M.V.; Bobić, J.D.; Grigalaitis, R.; Stojanović, B.D.; Banys, J. La-doped and La/Mn-co-doped barium titanate ceramics. Acta Phys. Pol. A 2013, 124, 155–160. [Google Scholar]
  14. Wodecka-Dus, B.; Plonska, M.; Czekaj, D. Synthesis, microstructure and the crystalline structure of the barium titanate ceramics doped with lanthanum. Arch. Metall. Mater. 2013, 58, 1305–1308. [Google Scholar] [CrossRef]
  15. Lines, M.E.; Glass, A.M. Principles and Applications of Ferroelectrics and Related Materials; Clarendon Press: Oxford, UK, 1977. [Google Scholar]
  16. Xu, J.; Zhai, J.; Yao, X. Structure and dielectric nonlinear characteristics of BaTiO3 thin films prepared by low temperature process. J. Alloys Compd. 2009, 467, 567–571. [Google Scholar] [CrossRef]
  17. Morrison, F.D.; Sinclair, D.C.; Skakle, J.M.; West, A.R. Novel doping mechanism for very-high-permittivity barium titanate ceramics. J. Am. Ceram. Soc. 1998, 81, 1957–1960. [Google Scholar] [CrossRef]
  18. Morrison, F.D.; Sinclair, D.C.; West, A.R. Electrical and structural characteristics of lanthanum-doped barium titanate ceramics. J. Appl. Phys. 1999, 86, 6355–6366. [Google Scholar] [CrossRef]
  19. Liu, X.; Hou, Y.; Song, B.; Cheng, H.; Fu, Y.; Zheng, M.; Zhu, M. Lead-free multilayer ceramic capacitors with ultra-wide temperature dielectric stability based on multifaceted modification. J. Eur. Ceram. Soc. 2022, 42, 973–980. [Google Scholar] [CrossRef]
  20. Wang, Y.; Miao, K.; Wang, W.; Qin, Y. Fabrication of lanthanum doped BaTiO3 fine-grained ceramics with a high dielectric constant and temperature-stable dielectric properties using hydro-phase method at atmospheric pressure. J. Eur. Ceram. Soc. 2017, 37, 2385–2390. [Google Scholar] [CrossRef]
  21. Agneni, A.; Paolozzi, A.; Sgubini, S. Piezoceramic devices modeled as mechanical systems in finite element codes. Cost 2001, 4, 591–599. [Google Scholar]
  22. Singh, D.; Dixit, A.; Dobal, P.S. Effect of Structural Changes on the Electrical Properties of Sol-gel derived Iron Doped Barium Ti-tanate. J. Phys. Conf. Ser. 2021, 2070, 012054–012062. [Google Scholar] [CrossRef]
  23. Staruch, M.; ElBidweihy, H.; Cain, M.G.; Thompson, P.; Lucas, C.A.; Finkel, P. Magnetic and multiferroic properties of dilute Fe-doped BaTiO3 crystals. APL Mater. 2020, 8, 031109. [Google Scholar] [CrossRef]
  24. Nayak, P.; Nayak, S.K. Doping effect of Fe, Co and W on the structural, electrical and magnetic properties of BaTiO3 ferroelectric ceramics. Solid State Commun. 2023, 371, 115272. [Google Scholar] [CrossRef]
  25. Hwang, S.-M.; Lim, J.-C.; Kim, S.-I.; Kim, J.-Y.; Hwang, J.; Lee, C.; Kwon, N.; Kim, I.; Lee, K.; Park, S.; et al. Impact of the change in charge compensation mechanism on the electrical, dielectric, and structural properties of La-doped BaTiO3 ceramics. J. Eur. Ceram. Soc. 2023. [Google Scholar] [CrossRef]
  26. Wodecka-Duś, B.; Czekaj, D. Electric Properties of La3+ Doped BaTiO3. Ceram. Ferroelectr. 2011, 418, 150–157. [Google Scholar] [CrossRef]
  27. Wodecka-Duś, B.; Lisińska-Czekaj, A.; Czekaj, D. Influence of lanthanum concentration on properties of BLT electroceramics. Key Eng. Mater. 2012, 512, 1308–1312. [Google Scholar] [CrossRef]
  28. Wodecka-Dus, B.; Adamczyk, M.; Dzik, J.; Osinska, K. The analysis of the electrical properties of BLT ceramics fabricated from sol-gel derived powders. Eur. Phys. J. B 2016, 89, 1–7. [Google Scholar] [CrossRef]
  29. Lin, F.; Jiang, D.; Ma, X.; Shi, W. Influence of doping concentration on room-temperature ferromagnetism for Fe-doped BaTiO3 ceramics. J. Magn. Magn. Mater. 2008, 320, 691–694. [Google Scholar] [CrossRef]
  30. Schwarzbach, J. Semiconductiving ceramic barium lanthanium titane doped with iron. Czechoslov. J. Phys. B 1968, 18, 1322–1334. [Google Scholar] [CrossRef]
  31. Zhe, T.Y.; Osman, R.A.M.; Idris, M.S. Crystal chemistry and electrical properties of La-doped BaTiO3. AIP Conf. Proc. 2021, 2347, 020008. [Google Scholar]
  32. Wodecka-Dus, B.; Adamczyk, M.; Osinska, K.; Płońska, M.; Czekaj, D. Dielectric properties of Ba1−xLaxTi1−x/4O3 ceramics with different La3+ content. In Advances in Science and Technology; Trans Tech Publications Ltd.: Zurich, Switzerland, 2013; Volume 77, pp. 35–40. [Google Scholar]
  33. Wodecka-Duś, B.; Adamczyk-Habrajska, M.; Goryczka, T.; Bochenek, D. Chemical and physical properties of the BLT4 ultra capacitor—A suitable material for ultracapacitors. Materials 2020, 13, 659. [Google Scholar] [CrossRef]
  34. Kumar, M.M.; Suresh, M.B.; Suryanarayana, S.V. Electrical and dielectric properties in double doped BaTiO3 showing relaxor behavior. J. Appl. Phys. 1999, 86, 1634–1637. [Google Scholar] [CrossRef]
  35. Sitko, D.; Garbarz-Glos, B.; Livinsh, M.; Bąk, W.; Antonova, M.; Kajtoch, C. Electrical Characterization of the Fe-Doped BT Ceramisc by an Impedance Spectroscopy. Ferroelectrics 2015, 486, 8–12. [Google Scholar] [CrossRef]
  36. Luo, B.; Wang, X.; Tian, E.; Song, H.; Zhao, Q.; Cai, Z.; Feng, W.; Li, L. Giant permittivity and low dielectric loss of Fe doped BaTiO3 ceramics: Experimental and first–Principles calculations. J. Eur. Ceram. Soc. 2018, 38, 1562–1568. [Google Scholar] [CrossRef]
  37. Chakraborty, T. Microscopic distribution of metal dopants and anion vacancies in Fe-doped BaTiO3-δsingle crystals. J. Phys. Condens. Matter. 2013, 25, 236002–236009. [Google Scholar] [CrossRef] [PubMed]
  38. Islama, A.; Momina, A.; Nesa, M. Effect of Fe doping on the structural, optical and electronic properties of BaTiO3: DFT based calculation. Chin. J. Phys. 2019, 60, 731–738. [Google Scholar] [CrossRef]
  39. Wodecka-Dus, B.; Goryczka, T.; Adamczyk-Habrajska, M.; Bara, M.; Dzik, J.; Szalbot, D. Dielectric and electrical properties of BLT ceramics modified by Fe ions. Materials 2020, 13, 5623. [Google Scholar] [CrossRef] [PubMed]
  40. Wodecka-Duś, B.; Kozielski, L.; Makowska, J.; Bara, M.; Adamczyk-Habrajska, M. Fe-Doped Barium Lanthanum Titanate as a Competitor to Other Lead-Free Piezoelectric Ceramics. Materials 2022, 15, 1089. [Google Scholar] [CrossRef] [PubMed]
  41. Baukamp, B.A. Electrochemical impedance spectroscopy in solid state ionics: Recent advances. Solid State Ion 2004, 169, 65–73. [Google Scholar] [CrossRef]
  42. Wodecka-Duś, B.; Kozielski, L.; Erhart, J.; Pawełczyk, M.; Radoszewska, D.; Adamczyk, M.; Bochenek, D. Investigation of La3+ doping effect on piezoelectric coefficients of BLT ceramics. Arch. Metall. Mater. 2017, 62, 691–696. [Google Scholar] [CrossRef]
  43. Hill, R.J.; Howard, C.J. Quantitative phase analysis from neutron powder diffraction data using the Rietveld method. J. App. Cryst. 1987, 20, 467–474. [Google Scholar] [CrossRef]
  44. Boukamp, B.A. A Linear Kronig-Kramers Transform Test for Immittance Data Validation. J. Electrochem. Soc. 1995, 142, 1885. [Google Scholar] [CrossRef]
  45. de Kronig, R.L. On the Theory of Dispersion of X-Rays. J. Opt. Soc. Am. 1926, 12, 547–556. [Google Scholar] [CrossRef]
  46. Macdonald, D.D.; Urquidi-Macdonald, M. Transformation of Constant Phase Impedances. J. Electrochem. Soc. 1990, 2, 515–517. [Google Scholar] [CrossRef]
  47. Bode, H.W. Network Analysis and Feedback Amplifier Design; Van Nostrand Reinhold: Princeton, NJ, USA, 1945. [Google Scholar]
  48. Rerak, M.; Makowska, J.; Osińska, K.; Zawada, A.; Adamczyk-Habrajska, M. The Effect of Pr Doping Contents on the Structural, Microstructure and Dielectric Properties of BaBi2Nb2O9 Aurivillius Ceramics. Materials 2022, 15, 5790. [Google Scholar] [CrossRef] [PubMed]
  49. Bochenek, D.; Dercz, G.; Chrobak, A. Electrical and magnetic properties of the BF–PFN solid solutions obtained by spark plasma sintering method. Mater. Sci. Eng. 2023, 295, 116625. [Google Scholar] [CrossRef]
  50. Cherakasova, N.; Veselov, S.; Bataev, A.; Kuzmin, R.; Stukacheva, N. Structure and mechanical properties of ceramic materials based on alumina and zirconia with strontium hexaaluminate additives. Mater. Chem. Phys. 2021, 259, 123938. [Google Scholar] [CrossRef]
  51. Płońska, M.; Plewa, J. Investigation of Praseodymium Ions Dopant on 9/65/35 PLZT Ceramics’ Behaviors, Prepared by the Gel-Combustion Route. Materials 2023, 16, 7498. [Google Scholar] [CrossRef]
  52. Zhang, B.; Wu, J.; Cheng, X.; Wang, X.; Xiao, D.; Zhu, J.; Wang, X.; Lou, X. Lead-free piezoelectric based on potassium-sodium niobiate with giant d33. ACS Appl. Mater. Interfaces 2013, 5, 7718–7725. [Google Scholar] [CrossRef]
  53. Pandey, S.; Kumar, D.; Parkash, O.; Pandey, L. Equivalent circuit models using CPE for impedance spectroscopy of electronic ceramics. Integr. Ferroelectr. 2017, 183, 141–162. [Google Scholar] [CrossRef]
  54. Rhouma, F.I.H.; Dhahri, A.; Dhahri, J.; Valente, M.A. Dielectric, modulus and impedance analysis of lead-free ceramics Ba0.8La0.133Ti1−xSnxO3 (x = 0.15 and 0.2). Appl. Phys. A 2012, 108, 593–600. [Google Scholar] [CrossRef]
  55. Shi, Y.; Pu, Y.; Cui, Y.; Luo, Y. Enhanced grain size effect on electrical characteristics of fine-grained BaTiO3 ceramics. J. Mater. Sci. Mater. Electron. 2017, 28, 13229–13235. [Google Scholar] [CrossRef]
  56. Dash, M.S.; Bera, J.; Ghosh, S. Effect of Porosity on Electrical Properties of Undoped and Lanthanum Doped BaTi0.6Zr0.4O3. In Proceedings of the 2007 IEEE International Conference on Solid Dielectrics, Winchester, UK, 8–13 July 2007. [Google Scholar]
  57. Sharma, P.; Kumar, P.; Kundu, R.S.; Juneja, J.K.; Ahlawat, N.; Punia, R. Structural and dielectric properties of substituted barium titanate ceramics for capacitor applications. Ceram. Int. 2015, 41, 13425–13432. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction of BLTF1, BLTF2, BLTF3, and BLTF4 ceramics.
Figure 1. X-ray diffraction of BLTF1, BLTF2, BLTF3, and BLTF4 ceramics.
Crystals 14 00131 g001
Figure 2. SEM images of the microstructure of the fractures of the obtained BLTF ceramics (magnification 5000×).
Figure 2. SEM images of the microstructure of the fractures of the obtained BLTF ceramics (magnification 5000×).
Crystals 14 00131 g002aCrystals 14 00131 g002b
Figure 3. Energy dispersive X-ray spectroscopy spectrum of BLTF ceramic samples.
Figure 3. Energy dispersive X-ray spectroscopy spectrum of BLTF ceramic samples.
Crystals 14 00131 g003
Figure 4. Frequency dependence of imaginary (Zim) and real (Zre) impedance component of BLTF ceramics for selected temperatures.
Figure 4. Frequency dependence of imaginary (Zim) and real (Zre) impedance component of BLTF ceramics for selected temperatures.
Crystals 14 00131 g004aCrystals 14 00131 g004b
Figure 5. Frequency dependence of residuals of the real and imaginary impedance of BLTF ceramics.
Figure 5. Frequency dependence of residuals of the real and imaginary impedance of BLTF ceramics.
Crystals 14 00131 g005
Figure 6. Figures of dependence of the imaginary part of the impedance (Zim) on the part of the real impedance (Zre) of the BLTF ceramics.
Figure 6. Figures of dependence of the imaginary part of the impedance (Zim) on the part of the real impedance (Zre) of the BLTF ceramics.
Crystals 14 00131 g006
Figure 7. Electric RC equivalent system with CPE phase-in element.
Figure 7. Electric RC equivalent system with CPE phase-in element.
Crystals 14 00131 g007
Figure 8. Experimental AC impedance spectrum in the complex plane (open circles) and modeled impedance spectrum using calculated values of RCPE equivalent circuit elements (solid red line) for BLTF1 ceramics at (a) 552 °C, (b) 502 °C, (c) 452 °C, and (d) 402 °C.
Figure 8. Experimental AC impedance spectrum in the complex plane (open circles) and modeled impedance spectrum using calculated values of RCPE equivalent circuit elements (solid red line) for BLTF1 ceramics at (a) 552 °C, (b) 502 °C, (c) 452 °C, and (d) 402 °C.
Crystals 14 00131 g008
Figure 9. Dependence of natural logarithm from the values of bulk resistances (R), obtained from the analysis of impedance spectra as a function temperature reversal for BLTF1 ceramics.
Figure 9. Dependence of natural logarithm from the values of bulk resistances (R), obtained from the analysis of impedance spectra as a function temperature reversal for BLTF1 ceramics.
Crystals 14 00131 g009
Figure 10. Electric equivalent system describing the electric properties of BLTF2, BLTF3, and BLTF4 ceramics.
Figure 10. Electric equivalent system describing the electric properties of BLTF2, BLTF3, and BLTF4 ceramics.
Crystals 14 00131 g010
Figure 11. Experimental AC impedance spectrum in complex plane (open circles) and modeled impedance spectrum using calculated values of RCPE-equivalent circuit elements (solid red line) for BLTF4 ceramics at (a) 550 °C, (b) 502 °C, (c) 452 °C, and (d) 402 °C.
Figure 11. Experimental AC impedance spectrum in complex plane (open circles) and modeled impedance spectrum using calculated values of RCPE-equivalent circuit elements (solid red line) for BLTF4 ceramics at (a) 550 °C, (b) 502 °C, (c) 452 °C, and (d) 402 °C.
Crystals 14 00131 g011
Figure 12. Dependence of natural logarithm from the values of grain (ac) and grain boundary resistances (df) (RG and RGB), obtained from the analysis of impedance spectra as a function of the reversal of temperature.
Figure 12. Dependence of natural logarithm from the values of grain (ac) and grain boundary resistances (df) (RG and RGB), obtained from the analysis of impedance spectra as a function of the reversal of temperature.
Crystals 14 00131 g012aCrystals 14 00131 g012b
Table 1. The real density (ρr) and open porosity (P) of BLTF ceramics.
Table 1. The real density (ρr) and open porosity (P) of BLTF ceramics.
Ceramicsρr
[g/cm3]
P
[%]
BLTF15.680.17
BLTF25.650.18
BLTF35.630.18
BLTF45.610.19
Table 2. The values of lattice parameters in BLTF composite.
Table 2. The values of lattice parameters in BLTF composite.
Compositea0 = b0
[nm]
c0
[nm]
δT
[nm]
V∙10−30
[m3]
BLTF10.39930.40321.009864.3
BLTF20.39920.40311.009864.2
BLTF30.39930.40311.009564.2
BLTF40.39930.40301.009364.2
Table 3. Parameters of the components of the applied electrical circuit for the impedance response of the ceramic BLTF1.
Table 3. Parameters of the components of the applied electrical circuit for the impedance response of the ceramic BLTF1.
BLTF1
T [°C]402452502552
RValue [Ω]1.626 × 106504,400171,00058,967
Relative error [Ω]29,4152158.4235.9498.275
Absolute error [%]1.8090.427910.137980.18911
CPE-TValue [F]4.314 × 10−96.427 × 10−98.676 × 10−91.0661 × 10−8
Relative error [F]5.7289 × 10−114.0278 × 10−112.9931 × 10−118.5592 × 10−11
Absolute error [%]1.32810.626680.344970.80285
CPE-PValue [a.u.]0.791310.768210.753570.74612
Relative error [a.u.]0.00110.00050.00030.0006
Absolute error [%]0.140.070.090.08
χ29.8 × 10−44.7 × 10−41.2 × 10−44.7 × 10−4
Table 4. Parameters of the components of the applied electrical circuit for the impedance response of the ceramic BLTF2, BLTF3, and BLTF4.
Table 4. Parameters of the components of the applied electrical circuit for the impedance response of the ceramic BLTF2, BLTF3, and BLTF4.
BLTF2
T [°C]552502452402
RGValue [Ω]9872086560918,602
Relative error [Ω]8.0723.39103.08587.2
Absolute error [%]0.821.121.833.15
CPEG-TValue [F]1.81 × 10−98.51 × 10−106.26 × 10−105.19 × 10−10
Relative error [F]1.11 × 10−104.99 × 10−114.47 × 10−115.11 × 10−11
Absolute error [%]6.135.867.109.84
CPEG-PValue [a.u.]0.8440.8960.9230.942
Relative error [a.u.]0.0040.0040.0050.008
Absolute error [%]0.460.440.580.88
RGBValue [Ω]829928,812109,150381,300
Relative error [Ω]34.30189.611614.415,156
Absolute error [%]0.410.661.483.97
CPEGB-TValue [F]5.11 × 10−85.45 × 10−84.57 × 10−82.52 × 10−8
Relative error [F]1.27 × 10−91.65 × 10−92.01 × 10−92.00 × 10−9
Absolute error [%]2.503.044.397.93
CPEGB-PValue [a.u.]0.7490.7230.7110.745
Relative error [a.u.]0.0020.0030.0050.009
Absolute error [%]0.310.430.691.33
χ20.000090.000410.00110.0041
BLTF3
RGValue [Ω]45979824209409
Relative error [Ω]0.701.293.5616.19
Absolute error [%]0.150.160.150.17
CPEG-TValue [F]2.14 × 10−101.87 × 10−101.59 × 10−101.54 × 10−10
Relative error [F]1.31 × 10−127.71 × 10−123.01 × 10−122.00 × 10−12
Absolute error [%]6.084.121.891.30
CPEG-PValue [a.u.]0.9450.9570.9730.985
Relative error [a.u.]0.0040.0030.0010.001
Absolute error [%]0.400.270.130.09
RGBValue [Ω]3285643620,64390,007
Relative error [Ω]5.9614.1654.61360.5
Absolute error [%]0.180.220.260.40
CPEGB-TValue [F]1.28 × 10−71.21 × 10−71.03 × 10−77.58 × 10−8
Relative error [F]1.29 × 10−91.31 × 10−91.06 × 10−98.48 × 10−9
Absolute error [%]1.0031.081.031.12
CPEGB-PValue [a.u.]0.8020.7980.7940.799
Relative error [a.u.]0.0010.0010.0010.001
Absolute error [%]0.120.130.140.18
χ28.33 × 10−50.000120.000140.00025
BLTF4
RGValue [Ω]523114229969559
Relative error [Ω]0.871.873.7921.46
Absolute error [%]0.170.160.130.22
CPEG-TValue [F]1.02 × 10−101.22 × 10−101.38 × 10−101.33 × 10−10
Relative error [F]6.7 × 10−124.25 × 10−122.31 × 10−122.12 × 10−12
Absolute error [%]6.53.471.721.59
CPEG-PValue [a.u.]0.9710.9650.9640.974
Relative error [a.u.]0.0040.0020.0010.001
Absolute error [%]0.420.260.120.11
RGBValue [Ω]588416,56650,118214,680
Relative error [Ω]5.7221.2175.611290
Absolute error [%]0.0970.130.150.80
CPEGB-TValue [F]4.14 × 10−83.5 × 10−82.82 × 10−82.19 × 10−8
Relative error [F]2.75 × 10−102.49 × 10−101.64 × 10−102.45 × 10−10
Absolute error [%]0.670.700.581.21
CPEGB-PValue [a.u.]0.7970.7960.8120.819
Relative error [a.u.]0.00050.0010.00060.001
Absolute error [%]0.0070.130.0730.157
χ24.11 × 10−57.43 × 10−55.99 × 10−50.0002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Adamczyk-Habrajska, M.; Wodecka-Dus, B.; Goryczka, T.; Makowska, J. Impedance Spectroscopy of Fe and La-Doped BaTiO3 Ceramics. Crystals 2024, 14, 131. https://doi.org/10.3390/cryst14020131

AMA Style

Adamczyk-Habrajska M, Wodecka-Dus B, Goryczka T, Makowska J. Impedance Spectroscopy of Fe and La-Doped BaTiO3 Ceramics. Crystals. 2024; 14(2):131. https://doi.org/10.3390/cryst14020131

Chicago/Turabian Style

Adamczyk-Habrajska, Malgorzata, Beata Wodecka-Dus, Tomasz Goryczka, and Jolanta Makowska. 2024. "Impedance Spectroscopy of Fe and La-Doped BaTiO3 Ceramics" Crystals 14, no. 2: 131. https://doi.org/10.3390/cryst14020131

APA Style

Adamczyk-Habrajska, M., Wodecka-Dus, B., Goryczka, T., & Makowska, J. (2024). Impedance Spectroscopy of Fe and La-Doped BaTiO3 Ceramics. Crystals, 14(2), 131. https://doi.org/10.3390/cryst14020131

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop