Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Investigation of the structure and dielectric properties of doped barium titanates

Mohamed M. Salem*a, Moustafa A. Darwish*a, Aseel M. Altarawneha, Yamen A. Alibwainiab, Ryad Ghazya, Osama M. Hemedaa, Di Zhouc, Ekaterina L. Trukhanovade, Alex V. Trukhanovde, Sergei V. Trukhanov*de and Maha Mostafaa
aPhysics Department, Faculty of Science, Tanta University, Al-Geish St., Tanta 31527, Egypt
bFaculty of Science and Information Technology, Jadara University, Irbid 21110, Jordan
cElectronic Materials Research Laboratory, Key Laboratory of the Ministry of Education, International Center for Dielectric Research, School of Electronic Science and Engineering, Xi'an Jiaotong University, Xi'an 710049, China
dSmart Sensors Laboratory, Department of Electronic Materials Technology, National University of Science and Technology MISiS, Moscow, 119049, Russia
eLaboratory of Magnetic Films Physics, SSPA “Scientific and Practical Materials Research Centre of NAS of Belarus”, 19, P. Brovki str., Minsk, 220072, Belarus. E-mail: s.v._trukhanov@gmail.com

Received 29th August 2023 , Accepted 3rd January 2024

First published on 22nd January 2024


Abstract

This work examined the influence of zirconium concentration on barium titanate (BZT) BaZrxTi1−xO3, with (x = 0, 0.15, 0.50, 0.75, and 1), produced by the tartrate precursor technique. The Fourier transform infrared (FTIR) spectra support the X-ray diffraction (XRD) results regarding formation of the perovskite structure. Grain size grows with Zr concentration, suggesting that the presence of Zr ions enlarges the grains. The transmission electron microscopy (TEM) images demonstrated that, due to their nano size, nanocrystallites are agglomerated in most images with irregular morphologies and average particle sizes from 20.75 nm to 63.75 nm. Increasing Zr content diminished the piezoelectric coefficient (d33) and the grain size. The value of d33 decreases by increasing Zr content, and there is an inverse relationship between grain size and d33. The remnant polarization of BZT increases with increasing Zr4+ content, which may be suitable for permanent memory device applications.


1. Introduction

Barium Zirconate Titanate (BZT) is a significant member of the BaTiO3 family. Much current technology relies on piezoelectric materials. Piezoelectric devices are used in computers; thermistors, capacitors, sensors, and piezoelectric transducers are used in security systems for various applications ranging from tiny speakers to medical ultrasounds.1–4

The BaZrxTi1−xO3 (BZT) are fascinating materials employed as dielectrics in industrial capacitor applications because of their high voltage resistance, dielectric constant, and composition-dependent Curie temperature (TC). Impurity doping in BZT is a popular method to enhance material performance.5–9 Several additives are often used with the barium titanate powder to improve performance and maintain appropriate control over the grain sizes and electrical characteristics. Doping BaTiO3 with additions like Sr2+, Ca2+, Sn4+, and Zr4+ allows for modifications. Because Zr4+ has better chemical stability than Ti4+, doping BaTiO3 with Zr4+ can enhance the dielectric and piezoelectric characteristics.

Barium titanate plays a significant role in the design and production of multiferroics. These materials simultaneously have at least two long-range orders in ordering the electronic or spin subsystems.10 To form multiferroic properties, either the formation of solid solutions11 or the receiving of composites12–14 is often used.

The current work aims to prepare BaZrxTi1−xO3 (x = 0, 0.15, 0.5, 0.75, and 1) nanopowders using tartrate precursor technique and studying the impact of the Zr4+ substitution ratio on the morphology, structure, piezoelectric, and ferroelectric characteristics of BZT was examined.

2. Materials and methods

As seen in Fig. 1a, in the synthesis of ferroelectric BaZrxTi1−xO3 (BZT) materials with varying Zr content (x = 0, 0.15, 0.50, 0.75, and 1) using the tartrate precursor method, a well-defined process comprising several essential steps was meticulously followed. It commenced with the preparation of mixtures of Ba–Zr–Ti solutions, wherein barium nitrate (Ba(NO3)2), zirconium oxide (ZrO2), and titanium nitrate (TiO2(NO3)2) served as the primary source materials. The precise stoichiometric quantities of these chemicals were carefully determined to correspond with the desired Zr content, as delineated in Table 1. Following the formulation of initial solutions, they underwent 15 minutes of stirring on a hot plate magnetic stirrer. Subsequently, an aqueous solution of tartaric acid (C4H6O6) was incorporated into the mixture, ensuring a homogeneous blend with continuous stirring. The subsequent phase involved the gradual evaporation of the solution through heating at a constant temperature of 80 °C while maintaining continuous stirring until reaching a state of dryness. The resulting dried powder was collected and further subjected to drying in a controlled environment at 100 °C. To foster the formation of the desired perovskite structure, the powder was then subjected to annealing at a high temperature of 1100 °C for 4 hours. The final stage of the synthesis process encompassed the pressing of tablets using a compressor, applying a substantial pressure of 5000 kg cm−2 while meticulously maintaining the thickness of the resulting tablets at d = 0.35 cm. Detailed records of the weights in grams of the individual chemical components used to synthesize BaZrxTi1−xO3 for each composition are presented in Table 1. This comprehensive description of the synthesis process and the precision in quantities employed underscores the methodological rigor and consistency in preparing ferroelectric BZT materials with varying Zr content.
image file: d3ra05885a-f1.tif
Fig. 1 (a) Samples preparation via the tartrate precursor method. (b) The experimental setup of d33 measurement.
Table 1 The weights in grams of BaZrxTi1−xO3
x Weight in grams
Ba(NO3)2 ZrO2 TiO2 (NO3)2 Molecular weight Total weight
0.0 26.132 0 20.38756 233.1922 46.51956
0.15 26.132 1.8483 17.329426 239.69575 45.309726
0.50 26.132 6.161 10.19378 254.8707 42.48678
0.75 26.132 9.2415 5.09689 265.70995 40.4703
1.00 26.132 12.322 0 276.5492 38.454


This process is a great way to get goods free of impurities at low temperatures. The samples were examined in the 2θ range between 20° and 80° using the X-ray diffraction (XRD) technique (Philips model PW-1729 diffractometer with Cu Kα radiation source with λ = 1.540598 Å).

The prepared samples' Fourier transform infrared (FTIR) spectra were obtained at room temperature (RT) using a PerkinElmer 1430 instrument in the wavenumber range from 200 cm−1 to 4000 cm−1. Transmission electron microscope (TEM) JEOL model 1010 was used to measure the particle size of the produced samples. A JEOL model JSM-5600 scanning electron microscope (SEM) was used to examine the sample morphology. The powders were pressed at 10 Ton into 1.3 cm-diameter discs form. After that, the discs (the pressed prepared samples) were polarized at 2 kV for two hours at RT for piezoelectric tests. By calculating the slope of the charge dependency of the applied stress, which was measured using the circuit in Fig. 1b, the piezoelectric coefficient (d33) was calculated. The Sawyer-Tower circuit has been used to evaluate the samples' ferroelectric hysteresis loop.

3. Results and discussion

3.1. X-ray diffraction and Rietveld refinement analysis

Fig. 2 shows the XRD patterns of the nanocrystalline of BZT annealed at 1100 °C for 4 h. The patterns indicated the tetragonal phase with space group (SG) P4mm.15 The diffraction peaks related to the tetragonal structure are (001), (110), (111), (200), (210), (211), (220), (310), (222) at 2θ = (21.2°, 30.10°, 37.1°, 43.1°, 48.5°, 53.5°, 62.6°, 71°, 75°, 79°) respectively. The sharp and defined peaks of the pattern indicate the high degree crystallinity and main intensity peak (110) increase by increasing Zr content.
image file: d3ra05885a-f2.tif
Fig. 2 XRD of BaZrxTi1−xO3 (x = 0, 0.15, 0.50, 0.75, 1).

The Zr ion is considered a non-catalyst agent. It abstracts the reaction process of the perovskite phase, which means that the reaction needs more time and energy for the phase, leading to the increased nanomaterial. The splitting of two diffraction peaks (002) at 2θ = 43.09° and (200) at 2θ = 44° confirm the presence of a tetragonal phase up to x = 0.75, whereas the sample x = 1 indicates the presence of cubic phase. The phase matching indicates that the tetragonal phase is the predominant one for the samples x = 0, 0.15, 0.5, and 0.75. At the same time, x = 1 has only a cubic phase.

The last sample has only cubic phase 100%, confirming our discussion. Because the electronic density of the Zr+4 ion (ionic radius of 0.72 Å) is higher than that of the Ti+4 ion (ionic radius of 0.60 Å), increasing the Zr+4 content results in an expansion of the unit cell volume, as shown by the primary diffraction peak (110) shifting to the lower angle side, suggesting an increase in the lattice parameter.16 The BaCO3 phase appears, and by increasing the Zr content at x = 0.15, 0.5, and 0.75, the BaCO3 phase decreases, meaning that annealing the prepared samples at 1100 °C is sufficient for forming perovskite phases with a low percentage of the undesired phases. The tolerance factor is given by the formula:17

 
image file: d3ra05885a-t1.tif(1)
where RA, RB, and RO are the ionic radius of A and B cations and oxygen ions, respectively. The tolerance factor value should be close to unity for a stable perovskite structure. Tolerance factor values of the samples were calculated using formula (1) and are given in Table 2, where their value range from 0.92 to 0.98, indicating the decreasing trend with increased Zr content due to the difference in ionic radius of Zr and Ti ions.18 The tolerance factor is reported to increase when a larger radius cation is substituted in the A-site. In contrast, substituted B-site may decrease the tolerance factor.4,5

Table 2 Tolerance factor and crystallite size of BaZrxTi1−xO3 (x = 0, 0.15, 0.50, 0.75, 1)
x Tolerance factor Crystallite size (D) (nm)
0 0.978 28.80
0.15 0.938 17.35
0.50 0.948 33.98
0.75 0.932 41.08
1 0.921 32.66


Scherer's equation is used to figure out the average crystallite size (D) of the perovskite structure:

 
image file: d3ra05885a-t2.tif(2)
where k = 0.89, θ is the position of the diffracted peak, λ is the wavelength of the X-ray source, and β is the full width at half maximum of the peak. The D of the BZT tetragonal phase was calculated and given in Table 2, which increased by increasing Zr content up to x = 0.75 and then decreased for the cubic phase x = 1.

Fig. 3 shows the Rietveld refinement of BaZrxTi1−xO3 (x = 0, 0.15, 0.5, 0.75, and 1) and indicates the presence of two tetragonal and orthorhombic phases for samples from x = 0 up to x = 0.75. The tetragonal phases with SG (P4mm) have a weight fraction of 48%, 49%, 82%, and 86%, respectively. The orthorhombic phases have a weight fraction of 52%, 51%, 18%, and 14% for all samples x = 0–0.75, respectively.


image file: d3ra05885a-f3.tif
Fig. 3 Rietveld-refined XRD patterns (left) and Rietveld refinement diffraction (right) of BaZrxTi1−xO3 (x = 0, 0.15, 0.5, 0.75, and 1).

The last sample (x = 1 or BaZrO3) has a cubic phase. The analyzed Rietveld refinement values of the tetragonal and orthorhombic phases are given in Table 3, with lattice parameters a, b, and c. It was noticed from Table 3 that for cubic system a = b = c = 4.18623 and tetragonal phase a = bc. The crystallite size for the two phases is also given in Table 3.

Table 3 Summary of Rietveld analyses of XRD data of BaZrxTi1−xO3 (x = 0, 0.15, 0.5, 0.75, and 1)
x Phase Structural parameter Crystallographic system SG Weight fraction (%) Density (g cm−3) Crystallite size (nm)
a (Å) b (Å) c (Å)
0 BaTiO3 4.185 4.185 4.195 Tetragonal P4mm 48.09 5.269 28.80
BaCO3 5.307 8.894 6.430 Orthorhombic Pmcn 51.91 4.138 44.56
0.15 BaZr0.15Ti0.85O3 4.193 4.193 4.201 Tetragonal P4mm 49.17 5.387 17.53
BaCO3 5.320 8.917 6.445 Orthorhombic Pmcn 50.83 4.287 29.83
0.50 BaZr0.5Ti0.5O3 4.178 4.178 4.192 Tetragonal P4mm 81.60 5.783 33.98
BaCO3 5.583 8.938 6.544 Orthorhombic Pmcn 18.41 4.031 30.09
0.75 BaZr0.75Ti0.25O3 4.182 4.182 4.197 Tetragonal P4mm 85.94 6.008 41.08
BaCO3 5.303 8.872 6.426 Orthorhombic Pmcn 14.06 4.335 35.27
1 BaZrO3 4.186 4.186 4.186 Cubic Pm[3 with combining macron]m 100 6.259 32.66


Fig. 3 shows the ball-and-stick model for the tetragonal and cubic phases, which are drawn using the crystallographic information with FullProf and Vesta software. It is shown in Fig. 3 that Ba-ions are in the corners of the unit cell with coordination (0,0,0). In contrast, Ti-ions are in the body center of the unit cell with coordination (0.5,0.5,0.5), as given in Table 4. The O-ions are in the phase center of the tetragonal unit cell. The Rietveld analyzer for sample x = 1 (100%) cubic phase with coordination of Ba and Zr ions is present in the body center.19

Table 4 Summary of the ball-and-stick model of BaZrxTi1−xO3 (x = 0, 0.15, 0.5, 0.75, and 1)
x Atom Atomic parameter
X Y Z
0 Ba 0 0 0
Ti 0.5 0.5 0.49
O 0.5 0 0.51
0.15 Ba 0 0 0
Ti 0.5 0.5 0.49
O 0.5 0 0.51
Zr 0.5 0.5 0.49
0.50 Ba 0 0 0
Ti 0.5 0.5 0.45
O 0.5 0 0.51
Zr 0.5 0.5 0.45
0.75 Ba 0 0 0
Ti 0.5 0.5 0.48
O 0.5 0 0.51
Zr 0.5 0.5 0.48
1 Ba 0 0 0
Zr 0.5 0.5 0.5
O 0.5 0 0


In contrast, the O-ions are at the mid-edge of the unit cell. The lattice parameters, crystallite size, and density of these phases are given in Table 4. The Rietveld refinement of the lattice constant agrees with the above discussion about the weight fraction of the present phases, where a = bc for the tetragonal and a = b = c for the cubic phase.

3.2. FTIR analysis

Fig. 4 shows the FTIR spectra for BaZrxTi1−xO3 where (x = 0, 0.15, 0.5, 0.75, and 1). The FTIR spectra showed two principal vibrational absorption bands around 350 cm−1 and 548 cm−1; other additional peaks appear at 3394 cm−1 due to the existence of O–H vibration of water adsorption by the sample, which has an asymmetric stretching mode which increases by increasing Zr content.
image file: d3ra05885a-f4.tif
Fig. 4 FTIR spectra of BaZrxTi1−xO3 (x = 0, 0.15, 0.5, 0.75, and 1).

Tartaric acid interacts with BaTiO3 during production, as evidenced by a peak at 1428 cm−1, attributed to an asymmetric vibration of the C–O bond.20 The first two absorption bands are associated with Ti–O stretching and bending vibrational modes. The peak at around 1143 cm−1 is attributed to the symmetrical CO2 and CH2 twisting mode. The absorption bands at 2464 cm−1 and 2927 cm−1 are related to the asymmetric and symmetric C–H stretching vibrations of the CH2 and CH3 groups, respectively.21 C–O deformation vibration peaks are around 1428 cm−1 and 1168 cm−1. The C–C stretching vibration is linked to the absorption band at 859 cm−1.

Table 5 shows the main frequency bands (ν1 and ν2) of BaZrxTi1−xO3 samples with different amounts of Zr and their force constants (F1 and F2). The peaks at 1762 cm−1 (C[double bond, length as m-dash]O stretching vibration), 1438 cm−1 (C–O asymmetric stretching vibration), 859 cm−1, and 693 cm−1 (C[double bond, length as m-dash]O bending vibration) all belong to the BaCO3 phase.22 As x increases from 0 to 0.75, the frequency of the absorption band at 691 cm−1 increases before dropping again. The initial absorption band, at 480 cm−1, has the same pattern of activity. By increasing the Zr content, shoulder ≅600 cm−1, we may explain this behavior as a modification in the bond length, given that the following equation23 was used to get the force constant for the A and B sites of the perovskite structure.

Table 5 The values of the main frequencies (ν1 and ν2), the force constants (F1 and F2), and the bond length of BaZrxTi1−xO3 (x = 0, 0.15, 0.5, 0.75, and 1)
x ν2 (cm−1) ν1 (cm−1) F2 (dyne per cm) ×104 F1 (dyne per cm) ×105 Bond length (Å)
0 351.03 545.81 8.7176 2.1076 2.0961
0.15 351.2 548.05 8.726 2.1249 2.0999
0.50 336.98 541.98 8.0337 2.0781 2.1013
0.75 352.96 553.55 8.8137 2.1678 2.0962
1 348.18 548.79 8.5766 2.1307 ____


3.3. SEM analysis

Fig. 5 displays the morphology of the BaZrxTi1−xO3 for (x = 0, 0.15, 0.5, 0.75 and 1). The presence of Zr ions increases the grain size, as seen by a correlation between grain size and Zr concentration. The upward trend could be attributable to the asymmetry in Zr and Ti ion distribution at the B-site. The SEM graph shows a clear separation in form, supported by the XRD results of the investigated samples, demonstrating an increase in the diffraction peak strength with increasing Zr concentration.24
image file: d3ra05885a-f5.tif
Fig. 5 SEM micrograph of BaZrxTi1−xO3 (x = 0, 0.15, 0.5, 0.75, and 1).

Table 6 shows that increasing grain size results in a higher density. Therefore, the presence of Zr concentration improves densification during sample preparation. Since the average grain size calculated from the SEM micrograph is greater than the crystallite size calculated using Scherer's equation from X-ray analysis, the grain is assumed to be occupancy of crystallites. Indicating that the tetragonal phase expands with both Zr content and grain size, the tetragonality factor c/a rises with increasing Zr concentration.25 Grains become bigger when Zr ions are replaced with Ti ions, indicating a more stable solid solution in the BaTi lattice.26

Table 6 Average grain size, density, crystallite size, and tetragonality factor of BaZrxTi1−xO3 (x = 0, 0.15, 0.5, 0.75, and 1)
x Grain size (μm) Density (g cm−3) Crystallite size (nm) Tertragonality factor (c/a)
0 1.54 5.269 28.80 1.0025
0.15 1.73 5.387 17.35 1.0018
0.50 1.78 5.783 33.98 1.0034
0.75 2.08 6.008 41.08 1.0035
1 2.80 6.259 32.66 1


3.4. TEM analysis

Fig. 6 showcases Transmission Electron Microscopy (TEM) images of Barium Zirconium Titanate (BaZrxTi1−xO3, abbreviated as BZT) synthesized via a tartrate precursor method. This methodology afforded us a series of compositions where zirconium content was systematically varied (x = 0, 0.15, 0.5, 0.75, and 1). In conjunction with the quantitative data presented in Table 7, these TEM images reveal a nuanced evolution of nanocrystallite sizes, ranging from 20.75 nm to 63.75 nm. A notable feature observed across multiple frames is the tendency of these nanocrystallites to agglomerate, resulting in irregular morphologies.15 This phenomenon indicates high surface energy and interaction potential among the nanocrystallites.
image file: d3ra05885a-f6.tif
Fig. 6 TEM images of BaZrxTi1−xO3 (x = 0, 0.15, 0.5, 0.75, and 1).
Table 7 Particle, crystallite, and grain size of BaZrxTi1−xO3 (x = 0, 0.15, 0.5, 0.75, and 1)
x Grain size from SEM (μm) Particle size from TEM (nm) Crystallite size from XRD (nm)
0 1.54 28 28.80
0.15 1.73 63.75 17.53
0.50 1.78 45.25 33.98
0.75 2.08 40 41.08
1 2.80 20.75 32.66


The average crystallite size demonstrates a complex dependence on the Zr content. In the absence of Zr (x = 0), the size is at its peak, suggesting an inherent stabilizing effect of the Ti-rich matrix. However, as the Zr content increases to x = 1, a diminishing trend in size is observed, likely due to the influence of the smaller ionic radius of Zr4+ compared to Ti4+. This size variation is aligned with a phase transition from a predominantly tetragonal structure (up to x = 0.75) to a primarily cubic phase in the final composition (x = 1). This transition is further corroborated by our phase analysis results.

Complementing the TEM analysis, Fig. 6 also presents Selected Area Electron Diffraction (SAED) patterns, crucial for confirming the crystalline nature of the synthesized BZT powders. These patterns reveal a perovskite-type tetragonal structure, which is in harmony with existing literature ref. 15, 27 and 28. The concentric halo rings observed in the electron diffraction patterns distinctly indicate the nanocrystalline nature of the material, with bright spots within these rings signifying high crystallinity. The appearance of these bright spots is a testament to the coherent scattering from well-ordered regions within the material.

Moreover, the SAED analysis provides an accurate delineation of crystalline lattice planes, notably (100), (111), and (210). Identifying these planes finds a direct correspondence in the X-ray diffraction (XRD) patterns, thereby establishing a vital link between SAED and XRD analyses.29 Our XRD investigations, detailed in Table 7, reveal that the sizes of the tetragonal crystallites are notably larger than those estimated using Scherer's formula. This discrepancy underscores the limitations of Scherer's formula, particularly in resolving smaller crystal sizes, as the lower detection limit of our XRD instrument stands at 20.75 nm.

Lastly, the efficacy of the tartrate precursor technique in yielding a nanocrystalline perovskite phase is amply demonstrated in the TEM images. These images show the crystallites' tetragonal shape and their transition to a cubic shape in the final composition (x = 1), mirroring the phase transition observed in XRD analysis. This alignment between morphological and crystallographic data underscores the intricate interplay between composition, crystal structure, and morphology in the BZT system.

3.5. Piezoelectric coefficient

Fig. 7 shows the impact of applied force on the produced samples' piezoelectric charge (Q33). It was found that the Q33 of the polarized sample declined as the applied stress increased, and we can compute the d33 in the unit (PC/N) from the slope of this line. The rotation of the 90° and 180° domains caused by an external electric field gives birth to the piezoelectric response.23,29,30 The d33 calculated values for all synthesized samples are shown in Table 8. It is noticed that the value of d33 decreases by increasing Zr content. The value of d33 is affected by the domain wall motion and the grain size, where the sample with a large grain size has lower piezoelectricity (d33), as given in Table 8.
image file: d3ra05885a-f7.tif
Fig. 7 The piezoelectric charge Q33 of the BZT samples for different Zr content.
Table 8 The values of d33, grain size, tetragonality factor, and crystalline size for different Zr content
Sample d33 (PC/N) Grain size (μm) Tetragonality factor (c/a) Crystalline size (nm) Reference
BaTiO3 10.08 1.54 1.0025 28.80 Present work
BaZr0.15Ti0.75O3 8.30 1.73 1.0018 17.53
BaZr0.50Ti0.50O3 7.94 1.78 1.0034 33.98
BaZr0.75Ti0.25O3 6.76 2.08 1.0035 41.08
BaZrO3 2.80 1 32.66
BaTiO3 31.1 30
BaTiO3 45 31
EVA/BaTiO3 0.85 32
BaTiO3/PVDF 18 33
BZT-BCT 10 34
BFZT-BT 19 35


Another aspect that influences the value of d33 is the tetragonality factor (c/a),23 which increases by increasing Zr content and is higher for sample x = 0.75 and equal to 1.047, as given in Table 8. Generally, the piezoelectric constant is correlated by the crystal structure, grain size, and density. The grain size has the most significant impact on the piezoelectric constant. The coupling effect between borders increases as grain size increases, producing a drop in domain mobility and polarization and, ultimately, a decrease in d33. It is essential to say the piezoelectric properties are strongly affected by the preparation method and the type of raw materials, which affect the grain size and the piezoelectric properties.31–33 As seen in Table 8, there is an inverse relationship between grain size and d33. The sample x = 1 deviates from the above discussion when the crystal structure is cubic. In the context of our research, we have clarified a fundamental relationship between grain size and the piezoelectric coefficient (d33). It is worth noting that grain size typically surpasses particle size in the hierarchical structure of the material. This phenomenon arises because an individual grain may encompass multiple agglomerated particles. Within this structural framework, an individual particle can exhibit the characteristics of a single, perfect crystal. In such cases, the dimensions of this particle align with what we call the crystallite size. However, a key consideration emerges when a particle presents certain low-angle defects, such as stacking faults or twin defects, which can reduce crystallite size. This parameter, reflecting the region characterized by perfect crystalline order, is routinely determined through X-ray diffraction (XRD) and consistently yields values smaller than those observed using electron microscopy.

A fundamental observation in our study is the discernment of an inversely proportional relationship between grain size and the piezoelectric coefficient (d33). Notably, this relationship is underscored by the empirical finding that larger grain sizes correspond to lower d33 values. This discernment emphasizes grain size's considerable impact on the material's piezoelectric behavior. In light of the pivotal role played by grain size in influencing d33, our investigation underscores the significance of understanding these structural parameters in the context of piezoelectric materials. Our result is in agreement with the previous work.34

3.6. The ferroelectric hysteresis-loop of BZT samples

Fig. 8 denotes the RT PE hysteresis loops for BZT samples. It is observed that the inclination of the hysteresis loop increases by increasing zirconium content up to sample x = 0.75 when sample x = 0 has a minimum inclination.36 The polarization increase for sample x = 0.75 due to the increase of 180° and 90° domains in the direction of the applied field.37 The remnant polarization was increased by increasing content up to x = 0.75, indicating that the Zr is a well-selected material for alignment for the domain of BZT samples,38 suggesting that the studied ferroelectric samples are relaxors ferroelectrics with soft hysteresis loops as shown in Table 9. At RT, all the present hysteresis loops for samples x = 0, 0.15, and 0.5 show relaxer ferroelectric behavior;39 these disordered hysteresis loops could be obeying the domain to is the relaxor material up to the applied field.40
image file: d3ra05885a-f8.tif
Fig. 8 The PE hysteresis loops of BaZrxTi1−xO3 samples where x = (0, 0.15, 0.5, 0.75, and 1).
Table 9 Ferroelectric parameters of BaZrxTi1−xO3 samples where x = (0, 0.15, 0.5, 0.75, and 1)
Composition (x) Ec (kV cm−1) Pr (μc cm−2) Ps (μc cm−2)
0 0.02 0.4 1
0.15 0.03 0.7 1.5
0.5 0.05 0.8 1.45
0.75 0.1 1 1.6
1 0.09 0.7 1.4


4. Conclusions

BaZrxTi1−xO3 samples with various Zr contents (x = 0, 0.15, 0.50, 0.75, and 1) were prepared using the tartrate precursor approach. According to the phase matching, the tetragonal phase predominates for samples with x = 0, 0.15, 0.5, and 0.75, whereas x = 1 only possesses a cubic phase. The crystallite size grew when the Zr content was raised to x = 0.75. When x = 1 was reached, the size was reduced. Ti ions are located in the cell's body center. In contrast, Ba-ions are at the corners of the unit cell with coordination (0,0,0) and (0.5,0.5,0.5), respectively. The sintering temperature of perovskite BZT powder is confirmed to be over 1100 °C by the compatibility of FTIR spectra with XRD investigation. Ti-ions replace Zr-ions to provide a more stable solid solution in the BaTi-lattice, which promotes grain enlargement. Zr is a well-selected material for alignment for the domain of BZT samples since the value of d33 declines with increasing Zr content and concentration up to x = 0.75, increasing residual polarization. The comprehensive approach, combining the tartrate precursor synthesis with extensive structural and functional analyses, has elucidated the complex interplay between Zr doping, crystal structure, grain size, and piezoelectric properties in BZT. This study advances our understanding of the material properties of doped barium titanates. It paves the way for developing novel ferroelectric materials with tailored properties for specific applications.

Author contributions

Conceptualization, M. M. S., O. M. H., M. A. D., E. L. T., A. V. T., S. V. T., and M. M.; methodology, M. M. S., A. M. A., Y. A. A., R. G., M. A. D., and M. M.; validation, M. M. S., A. M. A., Y. A. A., R. G., M. K., M. U. K., and O. M. H.; formal analysis, O. M. H., D. Z., M. A. D., E. L. T., M. K., M. U. K., A. V. T., and S. V. T.; investigation, A. M. A., Y. A. A., R. G., E. L. T., A. V. T., and S. V. T.; resources, O. M. H., Z. D., E. L. T., A. V. T., and S. V. T.; data curation, Z. D., M. A. D., E. L. T., A. V. T., and S. V. T.; writing—original draft preparation, M. M. S., E. L. T., A. V. T., S. V. T., and M. M.; writing—review and editing, M. M. S., E. L. T., A. V. T., S. V. T., and M. M.; visualization, A. M. A., Y. A. A., R. G., O. M. H., M. K., M. U. K., and M. A. D.; supervision, A. V. T., S. V. T., and M. M.; project administration, M. A. D., E. L. T., A. V. T., S. V. T., and M. M.; funding acquisition, M. A. D., E. L. T., M. K., M. U. K., A. V. T., and S. V. T. All authors have read and agreed to the published version of the manuscript.

Conflicts of interest

The authors declare no conflict of interest.

Acknowledgements

Alex V. Trukhanov thanks NUST MISIS for support within the framework of the “Priority 2030” (Smart Sensors Laboratory—project K6-2022-043).

References

  1. M. M. Hessien, N. El-Bagoury, M. H. H. Mahmoud and O. M. Hemeda, Synthesis and characterization of nanocrystalline barium-samarium titanate, High Temp. Mater. Processes, 2016, 35, 499–505,  DOI:10.1515/htmp-2015-0021.
  2. A. Prasatkhetragarn, T. Sareein, N. Triamnak and R. Yimnirun, Dielectric and ferroelectric properties of modified-BaTiO3 lead-free ceramics prepared by solid solution method, Ferroelectrics, 2022, 586, 224–241,  DOI:10.1080/00150193.2021.2014274.
  3. C. Filipič, Z. Kutnjak, R. Pirc, G. Canu and J. Petzelt, BaZr0.5Ti0.5O3: lead-free relaxor ferroelectric or dipolar glass, Phys. Rev. B, 2016, 93, 224105,  DOI:10.1103/PhysRevB.93.224105.
  4. T. Maiti, R. Guo and A. S. Bhalla, Evaluation of experimental resume of BaZrxTi1-xO3 with perspective to ferroelectric relaxor family: an overview, Ferroelectrics, 2011, 425, 4–26,  DOI:10.1080/00150193.2011.644168.
  5. M. Aamir, I. Bibi, S. Ata, K. Jilani, F. Majid, S. Kamal, N. Alwadai, M. A. S. Raza, M. Bashir, S. Iqbal, M. Aadil and M. Iqbal, Ferroelectric, dielectric, magnetic, structural and photocatalytic properties of Co and Fe doped LaCrO3 perovskite synthesized via micro-emulsion route, Ceram. Int., 2021, 47, 16696–16707,  DOI:10.1016/j.ceramint.2021.02.240.
  6. S. B. Li, C. B. Wang, X. Ji, Q. Shen and L. M. Zhang, Effect of composition fluctuation on structural and electrical properties of BZT-xBCT ceramics prepared by Plasma Activated Sintering, J. Eur. Ceram. Soc., 2017, 37, 2067–2072,  DOI:10.1016/j.jeurceramsoc.2016.12.043.
  7. Z. Liu, R. Yuan, D. Xue, W. Cao and T. Lookman, Origin of large electrostrain in Sn4+ doped Ba(Zr0.2Ti0.8)O3-x(Ba0.7Ca0.3)TiO3 ceramics, Acta Mater., 2018, 157, 155–164,  DOI:10.1016/j.actamat.2018.07.004.
  8. S. Ke, H. Fan, H. Huang, H. L. W. Chan and S. Yu, Dielectric dispersion behavior of Ba(ZrxTi1-x)O3 solid solutions with a quasiferroelectric state, J. Appl. Phys., 2008, 104, 034108,  DOI:10.1063/1.2964088.
  9. C. Zhang, Z. Zhou, Z. Tang, D. Ballo, C. Wang and G. Jian, Effects of scandium oxide on domain structure, dielectric and ferroelectric properties of barium zirconate titanate ceramics, J. Alloys Compd., 2022, 889, 161622,  DOI:10.1016/j.jallcom.2021.161622.
  10. V. Turchenko, A. S. Bondyakov, S. Trukhanov, I. Fina, V. V. Korovushkin, M. Balasoiu, S. Polosan, B. Bozzo, N. Lupu and A. Trukhanov, Microscopic mechanism of ferroelectric properties in barium hexaferrites, J. Alloys Compd., 2023, 931, 167433,  DOI:10.1016/j.jallcom.2022.167433.
  11. D. V. Karpinsky, M. V. Silibin, S. V. Trukhanov, A. V. Trukhanov, A. L. Zhaludkevich, S. I. Latushka, D. V. Zhaludkevich, V. A. Khomchenko, D. O. Alikin, A. S. Abramov, T. Maniecki, W. Maniukiewicz, M. Wolff, V. Heitmann and A. L. Kholkin, Peculiarities of the crystal structure evolution of BiFeO3–BaTiO3 ceramics across structural phase transitions, Nanomaterials, 2020, 10, 801,  DOI:10.3390/nano10040801.
  12. S. V. Trukhanov, A. V. Trukhanov, M. M. Salem, E. L. Trukhanova, L. V. Panina, V. G. Kostishyn, M. A. Darwish, An. V. Trukhanov, T. I. Zubar, D. I. Tishkevich, V. Sivakov, D. A. Vinnik, S. A. Gudkova and C. Singh, Preparation and investigation of structure, magnetic and dielectric properties of (BaFe11.9Al0.1O19)1-x - (BaTiO3)x bicomponent ceramics, Ceram. Int., 2018, 44, 21295–21302,  DOI:10.1016/j.ceramint.2018.08.180.
  13. M. M. Salem, L. V. Panina, E. L. Trukhanova, M. A. Darwish, A. T. Morchenko, T. I. Zubar, S. V. Trukhanov and A. V. Trukhanov, Structural, electric and magnetic properties of (BaFe11.9Al0.1O19)1-x - (BaTiO3)x composites, Composites, Part B, 2019, 174, 107054–107058,  DOI:10.1016/j.compositesb.2019.107054.
  14. R. E. El-Shater, A. S. Atlam, M. K. Elnimr, S. T. Assar, S. V. Trukhanov, D. I. Tishkevich, T. I. Zubar, A. V. Trukhanov, D. Zhou and M. A. Darwish, AC measurements, impedance spectroscopy analysis, and magnetic properties of Ni0.5Zn0.5Fe2O4/BaTiO3 multiferroic composites, Mater. Sci. Eng., B, 2022, 286, 116025,  DOI:10.1016/j.mseb.2022.116025.
  15. T. Badapanda, S. Sarangi, B. Behera, P. K. Sahoo, S. Anwar, T. P. Sinha, G. E. Luz, E. Longo and L. S. Cavalcante, Structural refinement, optical and ferroelectric properties of microcrystalline Ba(Zr0.05Ti0.95)O3 perovskite, Curr. Appl. Phys., 2014, 14, 708–715,  DOI:10.1016/j.cap.2014.02.015.
  16. C. Ostos, L. Mestres, M. L. Martínez-Sarrión, J. E. García, A. Albareda and R. Perez, Synthesis and characterization of A-site deficient rare-earth doped BaZrxTi1-xO3 perovskite-type compounds, Solid State Sci., 2009, 11, 1016–1022,  DOI:10.1016/j.solidstatesciences.2009.01.006.
  17. A. M. A. Henaish, O. M. Hemeda, A. M. Dorgham and M. A. Hamad, Characterization of excessive Sm3+ containing barium titanate prepared by tartrate precursor method, J. Mater. Res. Technol., 2020, 9, 15214–15221,  DOI:10.1016/j.jmrt.2020.10.015.
  18. N. Vonrüti and U. Aschauer, Band-gap engineering in AB(OxS1-x)3 perovskite oxysulfides: a route to strongly polar materials for photocatalytic water splitting, J. Mater. Chem. A, 2019, 7, 15741–15748,  10.1039/c9ta03116b.
  19. P. Sateesh, J. Omprakash, G. S. Kumar and G. Prasad, Studies of phase transition and impedance behavior of Ba(Zr,Ti)O3 ceramics, J. Adv. Dielectr., 2015, 5, 1550002,  DOI:10.1142/S2010135X15500022.
  20. N. Chakrabarti and H. S. Maiti, Chemical synthesis of barium zirconate titanate powder by an autocombustion technique, J. Mater. Chem., 1996, 6, 1169–1173,  10.1039/jm9960601169.
  21. O. M. Hemeda, M. E. A. Eid, T. Sharshar, H. M. Ellabany and A. M. A. Henaish, Synthesis of nanometer-sized PbZrxTi1-xO3 for gamma-ray attenuation, J. Phys. Chem. Solids, 2021, 148, 109688,  DOI:10.1016/j.jpcs.2020.109688.
  22. T. Thananatthanachon, Synthesis and characterization of a perovskite barium zirconate (BaZrO3): an experiment for an advanced inorganic chemistry laboratory, J. Chem. Educ., 2016, 93, 1120–1123,  DOI:10.1021/acs.jchemed.5b00924.
  23. O. M. Hemeda, B. I. Salem, H. Abdelfatah, G. Abdelsatar and M. Shihab, Dielectric and ferroelectric properties of barium zirconate titanate ceramics prepared by ceramic method, Phys. B, 2019, 574, 411680,  DOI:10.1016/j.physb.2019.411680.
  24. P. S. Aktaş, Structural investigation of barium zirconium titanate Ba(Zr0.5Ti0.5)O3 particles synthesized by high energy ball milling process, J. Chem. Sci., 2020, 132, 3–10,  DOI:10.1007/s12039-020-01837-7.
  25. M. Reda, S. I. El-Dek and M. M. Arman, Improvement of ferroelectric properties via Zr doping in barium titanate nanoparticles, J. Mater. Sci.: Mater. Electron., 2022, 33, 16753–16776,  DOI:10.1007/s10854-022-08541-x.
  26. N. Binhayeeniyi, P. Sukvisut, C. Thanachayanont and S. Muensit, Physical and electromechanical properties of barium zirconium titanate synthesized at low-sintering temperature, Mater. Lett., 2010, 64, 305–308,  DOI:10.1016/j.matlet.2009.10.069.
  27. H. H. Huang, H. H. Chiu, N. C. Wu and M. C. Wang, Tetragonality and properties of Ba(ZrxTi1-x)O3 ceramics determined using the Rietveld method, Metall. Mater. Trans. A, 2008, 39, 3276–3282,  DOI:10.1007/s11661-008-9622-2.
  28. O. M. Hemeda, B. I. Salem, H. Abdelfatah, G. Abdelsatar and M. Shihab, Dielectric and ferroelectric properties of barium zirconate titanate ceramics prepared by ceramic method, Phys. B, 2019, 574, 411680,  DOI:10.1016/j.physb.2019.411680.
  29. A. M. Henaish, M. Mostafa, I. Weinstein, O. Hemeda and B. Salem, Ferroelectric and dielectric properties of strontium titanate doped with barium, Magnetism, 2021, 1, 22–36,  DOI:10.3390/magnetism1010003.
  30. A. Tawfik, Elastic properties and sound wave velocity of PZT transducers doped with Ta and La, J. Am. Ceram. Soc., 1985, 68, 317–319,  DOI:10.1111/j.1151-2916.1985.tb10132.x.
  31. P. Zheng, J. L. Zhang, Y. Q. Tan and C. L. Wang, Grain-size effects on dielectric and piezoelectric properties of poled BaTiO3 ceramics, Acta Mater., 2012, 60, 5022–5030,  DOI:10.1016/j.actamat.2012.06.015.
  32. Y. Huan, X. Wang, J. Fang and L. Li, Grain size effects on piezoelectric properties and domain structure of BaTiO3 ceramics prepared by two-step sintering, J. Am. Ceram. Soc., 2013, 96, 3369–3371,  DOI:10.1111/jace.12601.
  33. V. R. Mudinepalli, L. Feng, W. C. Lin and B. S. Murty, Effect of grain size on dielectric and ferroelectric properties of nanostructured Ba0.8Sr0.2TiO3 ceramics, J. Adv. Ceram., 2015, 4, 46–53,  DOI:10.1007/s40145-015-0130-8.
  34. H. Shokrollahi, F. Salimi and A. Doostmohammadi, The fabrication and characterization of barium titanate/akermanite nano-bio-ceramic with a suitable piezoelectric coefficient for bone defect recovery, J. Mech. Behav. Biomed. Mater., 2017, 74, 365–370,  DOI:10.1016/j.jmbbm.2017.06.024.
  35. M. B. Ghasemian, Q. Lin, E. Adabifiroozjaei, F. Wang, D. Chu and D. Wang, Morphology control and large piezoresponse of hydrothermally synthesized lead-free piezoelectric (Bi0.5Na0.5)TiO3 nanofibres, RSC Adv., 2017, 7, 15020–15026,  10.1039/c7ra01293d.
  36. X. Li, M. Sun, X. Wei, C. Shan and Q. Chen, 1D piezoelectric material based nanogenerators: methods, materials and property optimization, Nanomaterials, 2018, 8, 188,  DOI:10.3390/nano8040188.
  37. O. Zahhaf, G. D’Ambrogio, A. Giunta, M. Q. Le, G. Rival, P. J. Cottinet and J. F. Capsal, Molten-State Dielectrophoretic Alignment of EVA/BaTiO3 Thermoplastic Composites: Enhancement of Piezo-Smart Sensor for Medical Application, Int. J. Mol. Sci., 2022, 23, 15745,  DOI:10.3390/ijms232415745.
  38. Z. Wang, L. Wang, Y. Meng, Y. Wen and J. Pei, Effects of Conductive Carbon Black on Thermal and Electrical Properties of Barium Titanate/Polyvinylidene Fluoride Composites for Road Application, J. Renewable Mater., 2023, 11, 2469–2489,  DOI:10.32604/jrm.2023.025497.
  39. S. A. Riquelme, K. Ramam and A. F. Jaramillo, Ceramics fillers enhancing effects on the dielectric properties of poly(vinylidene fluoride) matrix composites prepared by the torque rheometer method, Results Phys., 2019, 15, 102800,  DOI:10.1016/j.rinp.2019.102800.
  40. X. Shan, C. Zhou, Z. Cen, H. Yang, Q. Zhou and W. Li, Bi(Zn1/2Ti1/2)O3 modified BiFeO3-BaTiO3 lead-free piezoelectric ceramics with high temperature stability, Ceram. Int., 2013, 39, 6707–6712,  DOI:10.1016/j.ceramint.2013.01.110.

This journal is © The Royal Society of Chemistry 2024