Open Access Article
El Hassan Yahakoub
*a,
Khalid Lemrinia,
Talal Moudrikaha,
El Houcine Lahrarb,
Slimane Raissic,
Amine Bendahhou
a,
Ilyas Jalafia,
Fatima Chaoua,
Soufian El Barkanya and
Mohamed Abou-Salama
a
aDepartment of Chemistry, Laboratory of Molecular Chemistry, Materials and Environment (LCM2E)-The Multidisciplinary Faculty of Nador, University Mohamed Premier, B.P. 300, Selouane, Nador 62700, Morocco. E-mail: elhassan.yahakoub@ump.ac.ma
bLaboratory of Inorganic Materials for Sustainable Energy Technologies (LIMSET), University Mohammed VI Polytechnic, Benguerir, 43150, Morocco
cPSL University, Chimie ParisTech–CNRS, Institut de Recherche de Chimie Paris, 11 rue Pierre et Marie Curie, Paris 75005, France
First published on 10th December 2025
The solid solution Ba0.95Sm0.034Ti1−xZrxO3, with x = 0.01, 0.05, and 0.10, was synthesized via the solid-state reaction method at 1200 °C for 6 h. The obtained powders were first characterized by X-ray diffraction (XRD) to confirm the formation of the expected phase. All samples were found to exhibit a tetragonal structure with space group P4mm, as determined by Rietveld refinement of the XRD data. Increasing the zirconium content led to an expansion of the unit cell volume and a distortion of the [Ti/ZrO6] octahedra. Scanning electron microscopy (SEM) images of pellets sintered at 1300 °C for 6 h showed a denser microstructure with larger grains in Zr-rich ceramics. A significant modification of the optical bandgap energy was observed with increasing Zr content. Dielectric measurements showed that the Curie temperature gradually decreases with increasing zirconium content. The electrical properties were influenced by the contributions of grains and grain boundaries. The total electrical resistance (Rtot = Rg + Rgb) increased with the zirconium content. The analysis of the complex electrical modulus confirmed a non-Debye behavior for all the ceramics studied in this work.
The partial substitution of titanium by zirconium in barium titanate (BaTiO3) enhances the materials piezoelectric, ferroelectric, and relaxor properties compared to pure BaTiO3.8 The gradual introduction of Zr into the BaTiO3 structure leads to the formation of three distinct compositional domains within the BaTi1−xZrxO3 (BZT) solid solution, each characterized by specific dielectric responses and structural properties that depend on the zirconium content.9–11 For x values between 0 and 0.15, the material exhibits a significantly higher dielectric constant compared to unmodified BaTiO3. In the range 0.15 ≤ x ≤ 0.27, a diffuse phase transition is observed, indicating a gradual change in ferroelectric behavior. As the zirconium content increases further (0.27 ≤ x ≤ 0.40), the BZT compound develops a typical relaxor behavior, resulting from the continuous evolution of the diffuse phase transition.12 The polymorphic phase transitions identified in the BZT structure include successive transformations from rhombohedral to orthorhombic (T1), orthorhombic to tetragonal (T2), and finally from tetragonal to cubic at the curie temperature (TC). These transitions tend to converge as the Zr content increases, resulting in a decrease in the TC transition temperature, while the temperatures of the T1 and T2 transitions increase.13,14 To further enhance the materials properties, other strategies include substituting barium with alkaline earth elements of the same valence, such as Ca2+ and Sr2+,15–20 or with higher valence lanthanides such as La3+, Sm3+, Gd3+, Nd3+,Pr3+, Eu3+,21–23 as well as other cations like Y3+, Bi3+ or Sc3+.24–27 These approaches allow modulation of the electrical and dielectric properties while influencing the crystal structure and charge carrier mobility. Among these systems, the pseudo-binary system (1 − x)Ba(Zr0.2Ti0.8)O3–x(Ba0.7Ca0.3)TiO3 (BCZT) has attracted considerable attention due to the presence of a morphotropic phase boundary (MPB) and a “convergence region”, similar to those observed in the PZT and PMN-PT systems.28,29 This system exhibits outstanding energy storage properties and also shows great potential for cooling applications.30–32 Additionally, lanthanide modified BZT systems are attracting considerable interest because the introduction of higher valence ions such as La3+, Sm3+, Gd3+, Eu3+ creates a charge imbalance that leads to the formation of cation vacancies (at the Ba site) or oxygen vacancies in order to maintain electrostatic neutrality.27,33,34 These defects, together with the local lattice distortions induced by lanthanide ions, strongly affect charge-carrier mobility, electrical conductivity, and the dielectric and ferroelectric responses of the material. Altogether, these effects promote relaxor type behaviours and broaden the operational temperature range, which is particularly advantageous for high-performance applications.14,23,33,35–37 The incorporation of rare-earth elements (Re3+) at the A-site of the BZT system distorts the [Ti/ZrO6] octahedra, modifying the lattice symmetry and affecting the grain size, a key parameter for structural and electrical properties.14,33,35,38 This substitution also enhances the dielectric performance, as reported by Ostos et al., increasing the dielectric constant and stabilizing the structure for La3+, Nd3+, and Pr3+.39 Furthermore, the optical properties of the BZT system are also affected by this type of substitution, particularly when Ba is replaced by Gd, resulting in a reduction of the band gap energy (Egap).36,40
This study focuses on the synthesis and characterization of perovskite-structured materials, specifically the composition Ba0.95Sm0.034Ti1−xZrxO3, with the aim of optimizing the zirconium content while preserving ferroelectric behavior at room temperature. It seeks to address the following question: up to what zirconium content can titanium be substituted without compromising the room-temperature ferroelectricity, and how does this substitution affect the crystal structure, microstructure, as well as the dielectric, electrical, and optical properties of the synthesized material?
The crystal structure analysis of the ZrSm0.01, ZrSm0.05, and ZrSm0.10 powders by X-ray diffraction was performed using a D8 diffractometer (Bruker, Karlsruhe, Germany). The X-ray were generated by a copper anticathode (CuKα1 with λ = 1.5406 Å and CuKα2 with λ = 1.54439 Å), with a scanning speed of 0.012° per second at room temperature. Data acquisition was conducted at 40 kV and 40 mA. Rietveld refinement of the obtained data was performed using the Jana 2006 software.41 After X-ray diffraction analysis, the sintered pellets were carefully polished to obtain parallel surfaces, then annealed at 300 °C for 30 min. The morphology of the ceramics was analyzed using scanning electron microscopy (SEM) with a TESCAN VEGA III LM microscope, operating at an accelerating voltage of 10 KV. Optical properties of the material were evaluated using a UV-Visible spectrophotometer (UV-Shimadzu). The pellets were coated with a thin layer of silver paste, then thermally treated at 300 °C for 30 minutes to improve adhesion, before being cooled down to room temperature. This step was carried out prior to analysis by complex impedance spectroscopy. Impedance spectroscopy was performed using a BioLogic MTZ-35 analyzer coupled with a temperature-controlled oven. Measurements were carried out over a frequency range from 10 Hz to 1 MHz, applying an AC signal with an amplitude of 1 V.
![]() | ||
| Fig. 1 (a) X-ray diffractograms of the ZrSm0.01, ZrSm0.05, and ZrSm0.10 ceramics. (b) Magnified view of the 2θ region between 29.4° and 31.8°. | ||
The most intense peak in the XRD spectra is observed around 31° in each sample, as shown in Fig. 1(b). It is clear that the diffraction peaks tend to shift towards lower 2θ values as the zirconium content increases. According to Bragg's law,43 a decrease in the 2θ angle results in an increase in the crystal lattice volume. Indeed, the substitution of Ti4+ (RTi4+ = 0.605 Å) by Zr4+ (RZr4+ = 0.72 Å), due to the difference in ionic radii, induces a lattice expansion that may lead to an increase in the lattice parameters.44
To determine the structures or the phases present in the studied powders, peak deconvolution around 45° and 64°, as well as Rietveld refinement of the crystalline structures, were performed. For the reference structure of barium titanate, the diffraction peaks around 45° correspond to the (002)/(200)T reflections of the tetragonal phase, (200)/(022)O of the orthorhombic phase, (200)R of the rhombohedral phase, and (200)C of the cubic phase.45 The diffraction peaks located between 43° and 45° have been magnified and presented in Fig. 2(a)–(c) for the three compounds studied. The presence of a doublet in this region for each compound clearly suggests the presence of a tetragonal phase (P4mm),15 a peak corresponding to an orthorhombic phase (Amm2),46 or a mixture of both phases. Furthermore, the orthorhombic phase (Amm2) exhibits a triplet around 65°,10 corresponding to the (004), (040), and (222) reflections.46 On the other hand, the tetragonal phase (P4mm) shows a doublet in this region, corresponding to the (202) and (220) reflections. A zoom on the peaks located between 63° and 65° in the diffractograms is illustrated in Fig. 2(d)–(f). The presence of a doublet in this region, associated with the (202) and (220) reflections, strongly suggests that our samples are crystallized in the tetragonal phase (P4mm).
![]() | ||
| Fig. 2 (a), (b) and (c) show the diffraction peaks in the 45° region, while (d), (e), and (f) show the diffraction peaks in the 65° region for all compositions. | ||
To confirm the presence of the tetragonal phase in our ceramics, we performed a refinement using the Rietveld method, adopting an initial model based on a classic tetragonal perovskite structure with the formula Ba0.95Sr0.05TiO3, belonging to the P4mm space group, with lattice parameters a = 3.989 Å and c = 4.026 Å.47 In this model, Ba/Re atoms occupy the 1a site (0.5, 0.5, z), while Ti/Zr are located at the 1b site (0, 0, z). As established for the Ba0.95Sr0.05TiO3 structure, oxygen occupies two crystallographic ally distinct positions: the 1c site (0.5, 0, z) assigned to O1 and the 1d site (0, 0, z) assigned to O2.
Fig. 3, 4 and 5 show the Rietveld refinement spectra for all the studied compounds. The experimental XRD spectra and the calculated models exhibit good agreement for all compositions refined using the Rietveld method, confirming that all synthesized materials adopt a tetragonal perovskite structure with the P4mm space group. The refinement results, including the lattice parameters (a, c, V) and reliability factors (Rp, Rwp, GOF), are summarized in Table 1.
| Compounds | ZrSm0.01 | ZrSm0.05 | ZrSm0.10 |
|---|---|---|---|
| a (Å) | 4.0015 (3) | 4.00971 (5) | 4.01947 (17) |
| c (Å) | 4.0196 (5) | 4.02287 (8) | 4.02462 (14) |
| c/a | 1.0051 | 1.0033 | 1.0013 |
| V (Å3) | 64.0394 (11) | 64.6787 (18) | 65.1060 (4) |
| Rp (%) | 7.27 | 5.33 | 5.63 |
| Rwp (%) | 10.45 | 6.99 | 7.51 |
| Rexp (%) | 6.09 | 4.98 | 5.39 |
| GOF | 1.72 | 1.40 | 1.39 |
According to the results presented in this table, we can observe that the lattice parameters a and c increase with the rising Zr content. Consequently, the unit cell volume expands, which is consistent with the shift of X-ray diffraction peaks toward lower 2θ values. This volume increase can be attributed to the substitution of Ti4+ ions by Zr4+ ions. Additionally, the tetragonality ratio c/a decreases as the Zr4+ concentration increases, indicating a reduction in tetragonality. The (002) and (200) diffraction peaks, located between 44° and 45° (see Fig. 2(a)–(c)) for the three studied compositions, reveal that the increase in Zr4+ content and the corresponding decrease in Ti4+ content at the B site led to the merging of these peaks. This clearly demonstrates that the tetragonality of the crystal structure decreases with increasing Zr4+ concentration.
The quality of the refinement was assessed based on the reliability R-factors. All synthesized materials exhibited satisfactory refinement, as evidenced by the low values of Rp, Rwp, Rexp, and the goodness-of-fit (GOF), as presented in Table 1. The crystallographic parameters obtained, such as the atomic positions, Wyckoff positions, and the isotropic thermal displacement factor (Uiso) for each atom, derived from the refinement for all compositions, are compiled in Table 2.
| Compounds | Atoms | x | y | z | Uiso (Å) | Wyckoff position | Site occupancy |
|---|---|---|---|---|---|---|---|
| ZrSm0.01 | Ba/Sm | 0.5 | 0.5 | 0.5162 (12) | 0.0112 (5) | 1a | 0.95/0.034 |
| Ti/Zr | 0 | 0 | 0.022 (4) | 0.0123 (6) | 1b | 0.99/0.01 | |
| O1 | 0.5 | 0 | 0.010 (17) | 0.0160 (11) | 1c | 1 | |
| O2 | 0 | 0 | 0.492 (5) | 0.0160 (11) | 1d | 1 | |
| ZrSm0.05 | Ba/Sm | 0.5 | 0.5 | 0.5175 (14) | 0.0186 (14) | 1a | 0.95/0.034 |
| Ti/Zr | 0 | 0 | 0.016 (5) | 0.0170 (13) | 1b | 0.95/0.05 | |
| O1 | 0.5 | 0 | 0.042 (6) | 0.018 (18) | 1c | 1 | |
| O2 | 0 | 0 | 0.542 (7) | 0.018 (18) | 1d | 1 | |
| ZrSm0.1 | Ba/Sm | 0.5 | 0.5 | 0.5184 (7) | 0.0139 (5) | 1a | 0.95/0.034 |
| Ti/Zr | 0 | 0 | 0.041 (2) | 0.0118 (10) | 1b | 0.90/0.10 | |
| O1 | 0.5 | 0 | −0.014 (4) | 0.012 (2) | 1c | 1 | |
| O2 | 0 | 0 | 0.543 (9) | 0.012 (2) | 1d | 1 |
The distances between the atoms located in the A site (Ba/Sm), the B site (Ti/Zr), and the oxygens O1 and O2 are listed in Table 3. The analysis of these data shows that the displacement of (Ti/Zr) atoms within the B site along the c-axis decreases as the Zr concentration increases (see Fig. 6). This indicates a more pronounced distortion of the [Ti/ZrO6] octahedra in the Ba0.95Sm0.034Ti(1−x)ZrxO3 ceramic for SmZr0.10 compared to other compositions. The increase in this distortion in the Zr-rich compound could be associated with a reduction in tetragonality. This enhanced distortion is also a key factor in explaining the decrease in the Curie temperature (TC), as will be detailed in the section dedicated to dielectric properties.
| Bond distances (Å) | ZrSm0.01 | ZrSm0.05 | ZrSm0.10 |
|---|---|---|---|
| 4(Ti/Zr–O1) | 2.00131 (0) | 2.00753 (3) | 2.02459 (7) |
| 1(Ti/Zr–O2) | 2.13438 (3) | 2.11650 (5) | 2.01682 (9) |
| 1(Ti/Zr–O2) | 1.88758 (3) | 1.90637 (4) | 2.00265 (9) |
| 4(Ba/Ln–O1) | 2.82003 (2) | 2.77236 (4) | 2.75319 (8) |
| 4(Ba/Ln–O1) | 2.85349 (2) | 2.90918 (4) | 2.93785 (9) |
| 4(Ba/Ln–O2) | 2.83117 (3) | 2.83699 (4) | 2.84756 (10) |
![]() | ||
| Fig. 6 Representation of [Ti/ZrO6] octahedra for the ceramics (a) ZrSm0.01, (b) ZrSm0.05, and (c) ZrSm0.10. | ||
![]() | ||
| Fig. 7 (a), (b), and (c) are the SEM images, and (d), (e), and (f) are the histograms of the ZrSm0.01, ZrSm0.05, and ZrSm0.10. | ||
The chemical composition of the ceramics was determined using EDX spectroscopy. The electron beam causes ionization of the atoms' inner shells, leading to the emission of characteristic signals that are analyzed to identify the elemental composition of the ceramic under study. The EDX spectra presented in Fig. 8(a)–(c) show peaks corresponding to the elements Ba, Ti, Zr, Sm, and O, confirming the successful incorporation of all targeted ions into the crystalline matrix. The presence of carbon in the apparent composition of the samples is related to the sample holder used during the EDX analysis. The elemental composition of the samples was also examined by EDX (see the inset of Fig. 8(a)–(c)). Theoretically, for pure BaTiO3 ceramic, the atomic ratio Ba
:
Ti
:
O is close to 1
:
1
:
3.50 In our ZrSm0.01, ZrSm0.05, and ZrSm0.10 samples, the measured atomic percentages are very close to the theoretical values expected from the synthesis. Overall, the EDX results confirm the successful substitution of Ti4+ by Zr4+ in the Ba0.95Sm0.034Ti(1−x)ZrxO3 ceramics.
![]() | ||
| Fig. 9 (a) Absorbance spectra in the 250–600 nm range. (b–d) Plot of (αhν)2 as a function of photon energy hν for the ZrSm0.01, ZrSm0.05, and ZrSm0.10 ceramics. | ||
The optical bandgap energy (Egap) of the ZrSm0.01, ZrSm0.05, and ZrSm0.10 ceramics was determined using the Beer–Lambert law.43,52–54
![]() | (1) |
Table 5 presents the optical bandgap (Egap) values of the materials prepared in this work, along with those of various compounds reported in the literature. These data enable a comparison of the effect of substituting titanium (Ti4+) with zirconium (Zr4+) on the bandgap energy. The samples with the composition Ba0.95Sm0.034Ti(1−x)O3 (x ≤ 0.10), synthesized in the present work, exhibit optical bandgap (Egap) values ranging from 3.26 eV to 3.45 eV. These values are higher than those reported for compounds such as BaTiO3, BaTi0.95Zr0.05O3 and BaTi0.90Zr0.10O3, whose Egap values range between 3.09 eV and 3.20 eV. These results indicate that the moderate introduction of Zr4+ into the perovskite structure of BaTi(1−x)O3 (x ≤ 0.10) leads to a slight decrease in the bandgap energy. However, the co-substitution of Sm3+ at the Ba-site and Zr4+ at the Ti-site in compounds of the type Ba1−xSm2x/3Ti0.95Zr0.05O3 (x ≤ 0.03), leads to a gradual decrease in the bandgap energy. In general, the variation in bandgap energy is strongly dependent on the chemical composition of the materials. However, other factors can also influence this energy, including structural distortions and the formation of intermediate energy levels within the bandgap, which are associated with the introduction of trivalent ions.40,55 The optical band gap values obtained in this study (3.26 eV, 3.45 eV, and 3.41 eV) show that the Ba0.95Sm0.034Ti(1–x)ZrxO3 ceramics exhibit an optical gap compatible with all-dielectric photonic applications. Considering their ferroelectric nature and high-refractive-index characteristics, these materials could be integrated into metasurface or photonic crystal architectures to explore the generation of localized optical modes and the enhancement of emission or optical coupling. These perspectives therefore open the way for the potential use of the studied compositions in advanced optical devices and highly sensitive detection platforms.56
| Compounds | Egap | References |
|---|---|---|
| BaTiO3 | 3.10–3.20 | 57 |
| BaTi0.95Zr0.05O3 | 3.15 | 35 |
| BaTi0.90Zr0.10O3 | 3.09 | 58 |
| Ba0.99Sm0.0067Ti0.95Zr0.05O3 | 3.07 | 35 |
| Ba0.98Sm0.013Ti0.95Zr0.05O3 | 3.05 | 35 |
| Ba0.97Sm0.02Ti0.95Zr0.05O3 | 3.02 | 35 |
| Ba0.95Sm0.034Ti0.99Zr0.01O3 | 3.26 | Present work |
| Ba0.95Sm0.034Ti0.95Zr0.05O3 | 3.45 | Present work |
| Ba0.95Sm0.034Ti0.90Zr0.10O3 | 3.41 | Present work |
at TC for the ceramics with compositions ZrSm0.01, ZrSm0.05, and ZrSm0.10 are 7228, 4903, and 4611, respectively, at 1 kHz.
![]() | ||
| Fig. 10 Evolution of the relative permittivity for (a) ZrSm0.01, (b) ZrSm0.05, and (c) ZrSm0.10. (d) Comparison of the relative permittivity at 1 kHz for all prepared ceramics. | ||
These materials exhibit a clear dependence between the relative permittivity, the TC, and the increasing Zr content. The introduction of Zr reduces the maximum relative permittivity
, broadens the dielectric peak, and shifts the TC to lower values, likely due to lattice distortion induced by the substitution of Ti4+ with Zr4+.60 Fig. 10(d) shows the relative permittivity at 1 kHz for all synthesized ceramics as a function of temperature. The modification of the lattice parameter induces a shift of Ti4+/Zr4+ ions along the c-axis, thereby increasing the distortion of the [Ti4+/Zr4+O6] octahedra. This distortion weakens the coupling between the [Ti4+/Zr4+O6] octahedral groups, leading to a reduction in the TC and a decrease in the dielectric performance of the studied ceramics.
To analyze dielectric dispersion and the diffuse nature of phase transitions in ferroelectric materials, Uchino and Nomura introduced an empirical formula commonly referred to as the modified Curie–Weiss law.61,62 This expression is given as follows:
![]() | (2) |
represents the maximum relative permittivity at the temperature T′, A′ is the Curie–Weiss constant, the diffusivity coefficient γ equals 1 for a conventional ferroelectric and reaches 2 for a relaxor ferroelectric. When γ takes a value between 1 and 2, it signifies a partially diffuse phase transition.23,63,64 Fig. 11(a)–(c) shows the variation of
as a function of ln(T − Tm) at 1 MHz for all the prepared ceramics in the temperature range T > TC. The red line represents the fit to the modified Curie–Weiss law, while the scattered points correspond to the experimental data. A linear relationship is observed for all compositions, with γ values systematically calculated from the slope of each curve. The calculated γ values for ZrSm0.01, ZrSm0.05, and ZrSm0.10 ceramics are 1.259, 1.277, and 1.318, respectively. These γ values, which lie between 1 and 2, clearly indicate that these ceramics exhibit an incomplete diffuse phase transition. Furthermore, the gradual increase in γ with increasing Zr content suggests a continuous improvement in the relaxation behavior of the ceramics.60
Dielectric relaxation in these compositions occurs between frequencies of 103 Hz and 104 Hz. A shift towards higher frequencies is observed as the temperature increases. The dependence of relative permittivity on the composition of the prepared ceramics is shown in Fig. 12(d). It is clear that the dielectric permittivity decreases with increasing Zr content. Furthermore, the dielectric relaxation behavior becomes more pronounced as the Zr concentration increases. The enhancement of the relaxation behavior in dielectric ceramics is a crucial factor for optimizing their energy storage properties, as demonstrated in several studies.42,60,66
δ) with temperature at various frequencies for the ceramics ZrSm0.01, ZrSm0.05, and ZrSm0.10. Two distinct regions are observed, ranging from ambient temperature to 400 °C. The first region, from room temperature to 250 °C, exhibits significantly low values of tan(δ). The low tan(δ) values observed in this region result from the restricted electron diffusion at the grain boundaries.67 These grain boundaries, where the crystalline structures of adjacent grains meet, are typically sites of energy dissipation linked to electron mobility.38 In the ceramics ZrSm0.01, ZrSm0.05, and ZrSm0.10, the limited electron diffusion at the grain boundaries reduces energy loss, resulting in lower tan(δ) values. In the high-temperature range (250 °C to 400 °C), a significant increase in tan(δ) is observed. This behavior is mainly attributed to space charge polarization, which becomes prominent at elevated temperatures. Under these conditions, electrical conductivity dominates the dielectric response, while the influence of ferro-elastic domain walls remains minimal, leading to a rapid rise in tan(δ).68 The defining feature of ferroelectric materials is their ability to exhibit spontaneous polarization, which can be reversed under the influence of an external electric field. This behavior arises from the unit cell symmetry, which allows at least two equivalent directions for polarization. The material is divided into regions, called domains, where the polarization remains uniform, and these are separated by natural interfaces known as domain walls.69 The minimal influence of ferroelectric domain walls indicates that, at high temperatures, their contribution to tan(δ) is reduced. As a result, the value of tan(δ) increases with rising temperature.
A broad peak (anomaly) is found in the loss tangent graphs at high temperatures. This peak becomes more pronounced with increasing Zr concentration. The behavior observed in high-temperature regions is related to the dielectric relaxation process. This thermally activated relaxation mechanism is attributed to dipolar effects associated with the movement of charge carriers (electrons) within the grains.70
The fitting of the experimental data from the Nyquist diagrams using the appropriate equivalent electrical circuit with the MT-Lab software showed that the results are in good agreement with an equivalent electrical circuit consisting of two series-connected cells, corresponding to the grains and the grain boundary effects, respectively. Fig. 15 shows the fitting of the Nyquist diagram using the appropriate equivalent electrical model for ZrSm0.10. The various adjusted parameters are presented in Table 6.
| T (°C) | Rg (kΩ) | Qg (F s(α−1)) | αg | Rgb (kΩ) | Qgb (F s(α−1)) | αgb | |
|---|---|---|---|---|---|---|---|
| ZrSm0.01 | 300 | 720.074 | 6.64 × 10−9 | 0.7243 | 138.47 | 01.05 × 10−9 | 0.9774 |
| 320 | 422.832 | 6.31 × 10−9 | 0.7525 | 82.99 | 02.68 × 10−9 | 0.300 | |
| 340 | 290.585 | 7.42 × 10−9 | 0.7404 | 55.80 | 08.83 × 10−9 | 0.1617 | |
| 360 | 238.872 | 15.49 × 10−9 | 0.6582 | 31.75 | 21.22 × 10−9 | 0.1820 | |
| 380 | 216.350 | 35.39 × 10−9 | 0.5696 | 20.67 | 48.97 × 10−9 | 0.9999 | |
| 400 | 201.221 | 54.02 × 10−9 | 0.5250 | 17.64 | 16.82 × 10−9 | 0.9964 | |
| ZrSm0.05 | 300 | 1397 | 12.77 × 10−9 | 0.5556 | 79.183 | 3.052 × 10−9 | 0.6593 |
| 320 | 797.678 | 9.545 × 10−9 | 0.6182 | 49.351 | 14.14 × 10−9 | 0.4466 | |
| 340 | 472.167 | 7.863 × 10−9 | 0.6556 | 29.831 | 67.13 × 10−10 | 0.2492 | |
| 360 | 368.014 | 4.978 × 10−9 | 0.7020 | 17.605 | 48.98 × 10−10 | 0.2164 | |
| 380 | 320.112 | 5.298 × 10−9 | 0.6908 | 12.082 | 58.64 × 10−10 | 0.8656 | |
| 400 | 246.647 | 10.17 × 10−9 | 0.6346 | 9.879 | 45.45 × 10−10 | 0.9894 | |
| ZrSm0.10 | 300 | 1779 | 10.36 × 10−9 | 0.5975 | 52.308 | 06.78 × 10−9 | 0.5998 |
| 320 | 1020 | 8.879 × 10−9 | 0.6329 | 33.876 | 12.315 × 10−9 | 0.9594 | |
| 340 | 579.937 | 9.003 × 10−9 | 0.6158 | 16.176 | 14.62 × 10−9 | 0.5673 | |
| 360 | 381.627 | 6.112 × 10−9 | 0.6871 | 12.598 | 21.09 × 10−9 | 0.2962 | |
| 380 | 254.366 | 3.805 × 10−9 | 0.7676 | 10.322 | 25.44 × 10−9 | 0.9953 | |
| 400 | 191.644 | 17.25 × 10−9 | 0.5683 | 8.042 | 27.64 × 10−9 | 0.9996 |
The results presented in Table 6 show an increase in the values of Rg and a decrease in the values of Rgb with the increase in the amount of Zr. Additionally, it is observed that these values (Rg and Rgb) decrease with the rise in temperature for all ceramics, suggesting the thermal activation of conduction mechanisms in these materials. The values of Rg and Rgb were evaluated using the Arrhenius law to determine the activation energy associated with each.75,76
R = R0 exp(−Ea/kBT)
| (3) |
![]() | ||
| Fig. 16 Arrhenius plots of the data obtained for the ZrSm0.01, ZrSm0.05, and ZrSm0.10 ceramics: (a) for the grains, and (b) for the grain boundaries. | ||
| Compositions | Ea/eV (grain boundaries) | Ea/eV (grains) |
|---|---|---|
| ZrSm0.01 | 0.631 | 0.573 |
| ZrSm0.05 | 0.612 | 0.530 |
| ZrSm0.10 | 0.484 | 0.405 |
The activation energies of the grains range from 0.484 to 0.631 eV, while the activation energies of the grain boundaries range from 0.405 to 0.573 eV, for temperatures between 300 °C and 400 °C. It is clear that as the Zr content increases, the activation energy decreases.
The activation energy (Ea) represents the total energy required both to generate charge carriers and to enable their long-range migration, also known as the free-jump energy.77,78 In ferroelectric ceramics, this energy associated with small polarons typically ranges from 0.2 to 1.5 eV.79 It is influenced by the motion of domain walls, which depends on potential barriers at the grain boundaries that become more pronounced at elevated temperatures. These barriers are related to lattice distortions and grain size, both of which can limit the mobility of small polarons.80 The high activation energy observed for the low-zirconium-content compound is attributed to the low densification of these samples, as the presence of pores increases the energy required for charge carriers to perform their jumps.
![]() | ||
| Fig. 17 M″ as a function of frequency at various temperatures for (a) ZrSm0.01, (b) ZrSm0.05, and (c) ZrSm0.10. | ||
It is observed that, with increasing temperature, the relaxation peaks shift to higher frequencies. The presence of two peaks at distinct frequencies can be attributed to two different relaxation processes. At low frequency, the observed peak is associated with the long-range mobility of ions, which move from one site to another through hopping.81,82 At high frequency, however, the charge carriers are primarily confined within their potential wells, moving over short distances and undergoing only localized motions within these wells.83–88
The asymmetry of the relaxation peaks
reveals the presence of a stretching coefficient (β). This parameter allows characterization of the relaxation process type, whether Debye or non-Debye, and helps determine if the relaxation originates from the grains or the grain boundaries.84 To verify this, the stretching coefficient (β) was determined by fitting the imaginary part of the electric modulus (M″) using the Kohlrausch–Williams–Watts (KWW) function for a single M″ response, or the modified KWW function for a double response, according to Bergman's approach:89
![]() | (4) |
![]() | (5) |
represents the maximum value of
in the low-frequency region, while
corresponds to the maximum value in the high-frequency region. fgbmax is the peak frequency associated with the grain boundary anomaly, and fgmax corresponds to the grain-related anomaly. For ideal Debye-type relaxation, β = 1, whereas for non-Debye relaxation, β ranges between 0 and 1 (0 < β < 1). Fig. 18(a) and (b) shows the variation of β as a function of temperature, taking into account the contributions from both grains and grain boundaries. The β values, lying between 0 and 1, indicate a non-Debye behavior for all the ceramics studied. The relaxation times τg and τgb were analyzed using the Arrhenius law to determine the activation energies associated with the contributions of grains and grain boundaries.14
τM″ = τ0 exp(Ea/kBT)
| (6) |
| Compositions | Ea/eV (grains) | Ea/eV (grain boundary) |
|---|---|---|
| ZrSm0.01 | 0.635 | 0.736 |
| ZrSm0.05 | 0.609 | 0.689 |
| ZrSm0.10 | 0.498 | 0.498 |
Comparison of the activation energy values suggests that more energy is required for charge carriers to jump across the grain boundaries than within the grains. This can be explained by the fact that the movement of charge carriers inside the grains is confined to the oxygen octahedra of the unit cell, whereas at the grain boundaries, the diffusion or hopping of charge carriers requires more energy.88 It is important to note that a slight discrepancy was observed between the activation energies calculated from the Nyquist data (see Table 7) and those obtained from M″ data (see Table 8). This difference can be explained by the fact that Z″ primarily reflects the overall macroscopic behavior, while M″ is governed by the microscopic behavior of domains or grains within the sample.
| This journal is © The Royal Society of Chemistry 2025 |