Next Article in Journal
Artificial Neural Network for the Prediction of Fatigue Life of Microscale Single-Crystal Copper
Next Article in Special Issue
Effect of the Annealing Conditions on the Strain Responses of Lead-Free (Bi0.5Na0.5)TiO3–BaZrO3 Ferroelectric Thin Films
Previous Article in Journal
Magnetic Behavior of the Arrays of Iron Cylindrical Nanostructures: Atomistic Spin Model Simulations
Previous Article in Special Issue
Thermoelectrical Properties of ITO/Pt, In2O3/Pt and ITO/In2O3 Thermocouples Prepared with Magnetron Sputtering
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Synthesis and Domain Structures of Single-Crystal-Like Mesoscale BaTiO3 Plates

1
Electronic 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
2
School of Microelectronics, Xidian University, Xi’an 740071, China
*
Authors to whom correspondence should be addressed.
Crystals 2023, 13(3), 538; https://doi.org/10.3390/cryst13030538
Submission received: 27 February 2023 / Revised: 15 March 2023 / Accepted: 15 March 2023 / Published: 21 March 2023
(This article belongs to the Special Issue Lead-free Ferro-/Piezoelectric Ceramics and Thin Films)

Abstract

:
The (001) plate-like BaTiO3 piezoelectric micromaterials are synthesized by topochemical microcrystal conversion technique. BaTiO3 plates with a length of 2~10 μm and thickness of 0.5~1.3 μm are obtained. The dependence of morphology on synthesis conditions is discussed in detail. The crystal symmetry and multiscale domain structures of BaTiO3 plates are systematically investigated by various characterizations. X-ray diffraction (XRD) and Raman spectra analyses demonstrate the tetragonal symmetry of the (001) oriented BaTiO3 plates at room temperature. The domain configurations of the micron BaTiO3 are investigated with a polarized light microscope (PLM) and piezoresponse force microscopy (PFM). The single-crystal-like quality and uniformity are supported by PLM observations. More importantly, the classical 90° banded ferroelectric domains of ~125 nm width are observed for the first time in such BaTiO3 plates. The domain features in the mesoscale BaTiO3 plate are discussed and compared with its bulk counterparts. The results may provide insights into understanding and designing the mesoscale BaTiO3 functional materials.

1. Introduction

Barium titanate (BaTiO3) is a classical lead-free ferroelectric/piezoelectric material that was first discovered in the early 1940s [1]. BaTiO3 exhibits a cubic perovskite structure, which transforms into a tetragonal symmetry at the Curie point around 120 °C [2]. At room temperature, BaTiO3 was widely investigated in various material forms including single crystal and ceramics since its discovery, due to its excellent piezoelectric/ferroelectric/dielectric performances at room temperature.
Materials can display very distinct and interesting properties when the dimensions change from three-dimension bulk [3,4,5] to low-dimension nanoscale [6]. Generally, microscale and nanoscale refer to 1–100 μm and 1–100 nm, respectively, whereas mesoscale is somewhere in between. Significant advantages could be expected from those materials. Piezoelectric micro-/nano- materials and related devices have attracted more and more interest in recent years due to their small size and easy integration [7,8,9]. Two-dimensional ZnO platelet and the array assembled by ZnO platelets show great advantages as integrated piezotronic transistors [7]. Nanostructured piezoelectrics, for example, BaTiO3 were combined with polymer to form composite porous foam [8], which enhanced piezocatalysis due to the nanomaterial and the specific surface area. Moreover, nano- or mesoscale BaTiO3 materials can also be used as templates for piezoelectric ceramic texturing [10,11], which can enhance piezoelectric performance significantly. As a piezoelectric catalyst, nano- and micro-BaTiO3 particles were successfully utilized in wastewater degradation and H2 production [12,13].
It is believed that the ferroelectricity/piezoelectricity of BaTiO3 is critical to the desired functionalities, and how to synthesize good-quality mesoscale BaTiO3 with good performances is therefore in high demand. In contrast to bulk BaTiO3 and nanoscale BaTiO3 materials, mesoscale BaTiO3 is far less investigated. Single crystalline mesoscale BaTiO3 thin plates were first reported by Remeika in 1954 [14]. These clear, large size crystals have desired ferroelectric characteristics [15]. However, the potassium fluoride flux used during the growth of BaTiO3 single crystal is environmentally harmful and will cause the corrosion of furnace refractories. An eco-friendly and commonly used approach to obtain plate-like BaTiO3 is the topochemical microcrystal conversion (TMC) technique, which was first reported by Liu et al. [16]. In the TMC process, bismuth layer structure precursors transform to perovskite structure BaTiO3 in molten salt. There are two TMC methods reported on the synthesis of BaTiO3 plates, including the one-step TMC and the two-step TMC. The one-step TMC converts the Bi4Ti3O12 precursor to BaTiO3 directly. The two-step TMC converts Bi4Ti3O12 to BaBi4Ti4O15 first, and then obtains BaTiO3 from the BaBi4Ti4O15 precursor. With the topochemical reactions, the Ba2+ ions completely replace the Bi3+ ions in the precursor. Moreover, the plate-like morphology of the precursor, meanwhile, is maintained in the final products. BaTiO3 plates synthesized by the TMC method show a quadrate shape and a high aspect ratio. The obtained BaTiO3 particles are either single crystalline or polycrystalline, which is determined by the initial alignment after nucleation and the recrystallization process [17]. Much work has been focused on controlling the morphology of mesoscale BaTiO3 plates, which is mainly related to the following three factors. The first important factor is the choice of a suitable Aurivillius precursor. The morphology of the precursor directly affects the morphology of the final product during topochemical conversion. The morphology of BaTiO3 can be optimized by using a suitable precursor with a high aspect ratio, regular shape, and clean surface. Liu et al. utilized polycrystalline BaBi4Ti4O15 as a precursor for the preparation of BaTiO3 [16]. The plate-like and rectangular Aurivillius precursor of about 6~10 μm in length successfully converts to 5~10 μm perovskite BaTiO3 micrometer plates. Fan et al. found the exfoliation of BaTiO3 during TMC when they used Bi4Ti3O12 precursor or polycrystalline BaBi4Ti4O15 precursor [18]. The exfoliation resulted in irregular shape BaTiO3 plates and some fine particles among the plates. As a result, single-crystalline BaBi4Ti4O15 precursor [18] was reported to achieve uniform BaTiO3 plates. The size of BaTiO3 plates was also found to be closely related to the size of the precursor [19]. Micrometer and sub-micrometer BaTiO3 can be derived from a micro- or sub-micro-precursor, respectively. The second factor is the suitable molten salts. NaCl, KCl, and BaCl2 salts are the commonly used molten salts in the synthesis of BaTiO3 plates by the TMC method [16,17,18,19]. BaCl2-KCl molten salts systems are used in the transformation from Bi4Ti3O12 to BaBi4Ti4O15, because of their higher Ba2+ concentration to accelerate the growth of BaBi4Ti4O15 [16]. The NaCl-KCl molten salt system is used in the conversion from BaBi4Ti4O15 to BaTiO3. The last factor is the removal process of the salts and the byproducts. The resultant of the two-step TMC reaction includes BaTiO3 plate-like particles, salts, and residual Bi2O3. Salts can be easily removed after washing with deionized water several times. In order to remove the Bi2O3, an acid-washing process is needed with the reaction between Bi3+ and nitric acid (HNO3). Zuo et al. [20] compared two different washing methods. It was found to be better to first remove the salts by deionized water, and then eliminate the residual Bi3+ by HNO3.
On the one hand, potential applications are highly reliant on physical properties, such as piezoelectric and dielectric performances. On the other hand, the polar domain structures and corresponding dynamics can serve as an important extrinsic contribution to obtained electric properties. Therefore, the investigation of domain features is critical to the understanding of physical behaviors. However, the domain structure of BaTiO3 within the mesoscale is rarely reported. Wu et al. [13] reported PFM results of individual BaTiO3 plates. Though the amplitude butterfly loop and phase hysteresis loop were obtained along the out-of-plane direction, no obvious domains can be seen. Kržmanc et al. [19] found irregular domains in the PFM results of BaTiO3 plates, but the plates were stacked and the influence of polishing treatment cannot be ruled out.
In this work, we have optimized the synthesis of BaTiO3 plates by the TMC method, which guarantees the obtainment of single-crystal-like BaTiO3 materials in mesoscale. The crystal structure and microstructure of BaTiO3 plates are determined using X-ray diffraction (XRD), Raman spectra, and field-emission scanning electron microscope (FE-SEM). More importantly, direct observation of the domain structures in the BaTiO3 plates is performed by polarized light microscope (PLM) and piezoresponse force microscopy (PFM). The good quality sample enables the observation of clear domain configurations. The domain features of BaTiO3 materials in mesoscale are discussed by comparing them with their bulk counterparts. The results may provide new insights into understanding and designing the mesoscale BaTiO3 functional materials.

2. Materials and Methods

2.1. The Synthesis of the Precursors and the BaTiO3 Plates

The BaTiO3 plates were prepared by the TMC method in three steps. First, a stoichiometric amount of Bi2O3 (99%) and TiO2 (98%) with respect to the Bi4Ti3O12 precursor were mixed by ball milling in ethanol for 15 h. The mixture was then dried and calcined under 1080~1160 °C using an equal weight of molten salts. The molar ratio of NaCl (99.5%) and KCl (99.5%) salts was 1:1. Subsequently, BaBi4Ti4O15 precursor particles were synthesized in 1:1 mol BaCl∙2H2O (99.5%) and KCl (99.5%) salts. The salts, BaCO3 (99%) and TiO2 (98%), were ball milled for 15 h. Then, the Bi4Ti3O12 particles obtained in the first step were added and magnetically stirred with ethanol for 8 h. After drying, the mixture was calcined at 1040~1080 °C. An excess of BaCO3 and TiO2, equivalent to 10 mol% were added prior to the ball-milling process. Finally, BaTiO3 was synthesized from the mixture of BaBi4Ti4O15 and BaCO3 (99%) with an equal weight of NaCl (99.5%) and KCl (99.5%) salts. The salts and BaCO3 were ball milled, and then magnetic stirred with BaBi4Ti4O15 particles for 8 h in ethanol. The mixture was dried at 90 °C. The calcined temperature was 925~975 °C. The heating rate was 4 °C/min throughout the calcination process. The salts of each step were removed by washing them several times in deionized water. Especially, in the third step, an acid-washing process was needed to remove the Bi2O3 byproduct. The calcined powder of the third step was washed in HNO3 solution for 2 h, and then washed in deionized water several times.

2.2. Characterization

The phase structure was determined using a high-resolution X-ray diffractometer (XRD, PANalytical Empyrean, PANalytical, Almelo, the Netherlands) with Cu Kα1 radiation. The Raman spectra were collected by Raman spectrometer (RTS2-501-SMSXS, Zolix, Beijing, China) with a 532 nm laser at room temperature. The morphology of the BaTiO3 and the precursor particles were examined using field-emission scanning electron microscopy (FE-SEM, FEI Quanta 250 FEG, FEI, Minato, Japan). Energy dispersive spectroscopic (EDS) mapping was carried out on the BaTiO3 plates. A polarized light microscope (PLM, Olympus BX-51, Olympus, Tokyo, Japan) was used to investigate the extinction angle and the domain structure of the BaTiO3. The domain structures of the BaTiO3 plates were characterized by a piezoresponse force microscope (PFM, Dimension ICON, Bruker, Billerica, MA, USA).

3. Results and Discussion

The structure of Aurivilius compounds can be described as the intergrowth of (Bi2O2)2+ layers and (An−1BnO3n+1)2− layers, where n (n = 1–6) is the number of octahedral layers between two neighboring bismuth oxides layers [21]. Aurivilius-structured Bi4Ti3O12 was chosen as the precursor, and was obtained by the reaction between Bi2O3 and TiO2 according to Equation (1). In the second step of the TMC process, Ba2+ partially substituted Bi3+ in the Bi4Ti3O12 precursor and formed BaBi4Ti4O15 (Equation (2)). In the final step, Ba2+ substituted all the residual Bi3+ in the BaBi4Ti4O15, resulting in the formation of the final product BaTiO3 (Equation (3)).
2 Bi 2 O 3 + 3 TiO 2 Bi 4 Ti 3 O 12
Bi 4 Ti 3 O 12 + BaCO 3 + 4 TiO 2 BaBi 4 Ti 4 O 15 + CO 2
BaBi 4 Ti 4 O 15 + BaCO 3 2 BaTiO 3 + 2 Bi 2 O 3 + CO 2
Figure 1 shows the FE-SEM images of the Bi4Ti3O12, BaBi4Ti4O15 precursor particles and the BaTiO3 plates synthesized at different temperatures. The results revealed that all the particles possess a plate-like shape. The morphology of the Bi4Ti3O12 particles is shown in Figure 1a–c.
BaBi4Ti4O15 nucleation initiates on the Bi4Ti3O12 precursor and further grows by Ostwald ripening during the heating and holding stages. Generally, smaller particle size Bi4Ti3O12 leads to smaller BaBi4Ti4O15 after the topochemical reaction. Notably, the Bi4Ti3O12 reaches a larger particle size at higher temperatures but shows obvious agglomeration in Figure 1c. The thickness of Bi4Ti3O12 is about 0.2~1 μm. The length distributions are shown in Figure 2a–c. The average length of the Bi4Ti3O12 synthesized at 1080 °C, 1120 °C, and 1160 °C are about 4.43 μm, 7.54 μm, and 12.51 μm, respectively. The Bi4Ti3O12 obtained at 1080 °C in Figure 1a is too small and not suitable as a precursor. Bi4Ti3O12 particles that grew at 1120 °C were selected as the precursor for the second step. BaBi4Ti4O15 obtained at 1040 °C, 1060 °C, and 1080 °C are 6.23 μm, 7.73 μm, and 8.91 μm in length, respectively. The distribution curves show better uniformity at 1040 °C and 1060 °C temperature conditions, according to Figure 2d–f.
The FE-SEM images in Figure 1g–i show the plate-like shaped BaTiO3 plates converted from the optimized BaBi4Ti4O15 of the second step. The morphology of BaTiO3 synthesized at 925 °C and 950 °C are similar. Figure 1g shows the BaTiO3 synthesized at 925 °C with an average particle size of 6.18 μm. The particles possess suitable size in the range of 3–10 μm, but are not uniformly distributed. Moreover, the BaTiO3 in Figure 1i contains many irregular shape particles with poor uniformity, which is also evidenced by Figure 2i. The BaTiO3 plates obtained at 950 °C in Figure 1h are approximately 6.79 μm in length with a relatively regular square shape. The particles in Figure 1h are about 2~10 μm in length and about 0.5~1.3 μm in thickness. As shown in Figure 1h and Figure 2h, the plates show good uniformity without obvious agglomeration. The products with the optimized condition of the three steps show no obvious changes in length between the precursors and BaTiO3 plates, which means minimized exfoliation. These mesoscale BaTiO3 particles exhibit a clean surface, and no obvious crack or hole was found. Note the obtained large anisotropic morphology and high aspect ratio are important for BaTiO3 as templates in ceramic texturing applications. Moreover, the large specific surface area is also in favor of potential piezocatalysis applications.
Figure 3a shows the XRD patterns of the BaTiO3 plates synthesized at 950 °C. Phase-pure BaTiO3 particles were obtained. The diffraction pattern in Figure 3a can be well indexed with a standard PDF card (No. 81-2202) with a P4mm space group. Moreover, the lattice parameter of the BaTiO3 plates is calculated to be a ≈ 4.00 Å and c ≈ 4.04 Å, which agree well with the reported value in standard PDF card, i.e., a = b = 3.99 Å, c = 4.04 Å. Figure 3b displays the Raman spectra measured in the frequency range of 100 cm−1 − 900 cm−1. The inset is the photograph of the measured plate on a platinum substrate. The Raman spectra show peaks at 183 cm−1, 301 cm−1, 510 cm−1, and 721 cm−1, which is consistent with the literature [22,23,24,25]. The peak at ~180 cm−1 is related to the decoupling of the A1(TO) phonons [19]. A weak peak at 301 cm−1 suggests the [B1, E(TO + LO)] mode related to the TiO6 octahedra [13,26]. The peaks at 510 cm−1 and 721 cm−1 correspond to the Raman active mode of [E(TO),A1(TO)] and [E(LO),A1(LO)], respectively.
Figure 4 exhibits the EDS mapping results on the surface of BaTiO3 plates. The measured single plate is square in shape with a length of about 7 μm as shown in Figure 4a. The dotted lines in Figure 4b–d mark the boundary of the plate. Elements distribution of Ba, Ti, and O elements are shown in red, green, and yellow colors, respectively. Generally, Ba, Ti, and O elements are homogeneously distributed in the plates, suggesting good uniformity of the sample. Note in Figure 4d, a lower content of O elements was detected beyond the region of the plate, which may originate from the surrounding circumstance.
PLM images of the (001) face of BaTiO3 plates are depicted in Figure 5. The white arrow marks the direction of [100] along the edge of the BaTiO3 plate. The images with crossed polarizer (P) and analyzer (A) are shown in Figure 5b–d at different intersection angles of 0°, 45°, and 90° between the analyzer and [100] direction. As shown in Figure 5b–d, the mesoscale BaTiO3 plate reveals an extinction angle of 90 degrees, which is consistent with the tetragonal symmetry evidenced by XRD characterization in Figure 3. Moreover, the complete extinction characteristics observed not only indicate the good uniformity of symmetry in the sample, but also suggest the unified orientation, i.e., the geometry of the template is coincident with the microscale perovskite unit cell. This also supports the single-crystal-like quality of the sample. The sample becomes color at 45° (the brightest position), reflecting the effect of a-domains in tetragonal BaTiO3. However, no obvious regular domain patterns can be clearly seen. One possible reason is that the domain size is too small to be detected, as PLM cannot detect sub-micrometer scale domains. Another reason could be the domain overlapping, which may also make the PLM observation difficult.
To further investigate the local domain structure in the micro-scale BaTiO3 plates, PFM was performed on an individual plate. It can be seen that the out-of-plane as well as the in-plane amplitude images exhibit distinct contrast of about 180°, which is also revealed in the corresponding phase images. Figure 6a shows an out-of-plane PFM image recorded on the (001) surface of the plate, whereas the normal edge of the plate is oriented parallel to <001>-axis or <010>-axis. Figure 6c–f are the enlarged view of the region marked in Figure 6a. The roughness of this enlarged region extracted from the surface morphology image is ~43.5 nm. The rough surface is consistent with the FE-SEM results observed in Figure 1 and Figure 4, which may be caused by the variation in growth rate in the different regions of the plate. Out-of-plane and in-plane images in the region of 6 μm × 6 μm area were recorded. The bright and dark stripes show the in-plane polarization with opposite directions along the <100>-axis (or <010>-axis), respectively. In the out-of-plane images, the bright and dark stripes indicate polarization with opposite directions along the <001>-axis. The BaTiO3 plate shows a group of regular 90° ferroelectric banded domains with an average domain width of ~125 nm. In Figure 6b, a 3D schematic domain structure is sketched according to the possible orientation of the domains in the T phase (001) plate. Arrows indicate the orientation of spontaneous polarization (Ps). In the a1/a2-domain, the Ps is parallel with the surface of the plate, marked as P1. In the c-domain, the Ps is perpendicular to the plate surface, marked as P2. The green stripes mark the c-domains while pink stripes represent the a-domains as illustrated in Figure 6b. The walls between 90° a/c domains intersect with the (001) plane by an angle of 45°. In the in-plane images, the bright and dark stripes correspond to P1 and P2, respectively. Meanwhile, in the out-of-plane images, the P1 and P2 regions also show distinct contrast. However, the bright and dark stripes indicate the P2 and P1, which are opposite to the in-plane signals. In previous studies, only irregularly shaped domain patterns were reported in mesoscale BaTiO3 plates [19]. The results in Figure 6 show the first observation of typical 90° banded domains in such BaTiO3 plates in mesoscale.
The features of the domain in mesoscale BaTiO3 plates compared with bulk BaTiO3 are worth further discussion. Different from the randomly oriented grains in BaTiO3 ceramics, the BaTiO3 plates possess a preferred orientation and much smaller sample thickness. The dimensions have a notable influence on the domain size. Though BaTiO3 in the T phase is known to display typical 180° and 90° ferroelectric domain features [27], domain size can vary significantly with different microstructures. In the case of ceramics with different grain sizes, 90° domains of sub-micrometer scale were reported in the BaTiO3 ceramics with 8.61 μm grains, whereas nanodomains (10~30 nm) were found in the ceramics of grain size <1 μm [28]. Also in BaTiO3 single crystals, multiscale domain patterns covering a wide range from nanometers to several tenths of micrometers [29,30,31] are observed. Kittel’s law shows the relation between the domain width w and the sample thickness t in ferroelectrics, ferromagnetics, and ferroelastics [32,33]. The domain width decreases with the decreasing thickness, i.e., w2 = kt. The theoretical constant of proportionality k in this relation is reported to be 8.5 × 10−9 m for BaTiO3 single crystals [34]. Adopted with the same parameter k, the calculated domain width w is in the range of 65~105 nm based on experimental thickness data. This is in the same order as our measured results ~125 nm, indicating the size dependence in mesoscale BaTiO3 plate generally coincides with bulk single crystals.

4. Conclusions

The mesoscale plate-like BaTiO3 was synthesized by the TMC technique in three steps. The Bi4Ti3O12, BaBi4Ti4O15 precursor, and the final product BaTiO3 prepared at different processing temperatures in each step were compared. The length distribution of the Bi4Ti3O12, BaBi4Ti4O15, and BaTiO3 particles were analyzed based on the FE-SEM results. The dependence of morphology on synthesis conditions were discussed in detail. Micron-size plate-like BaTiO3 plates with near square shape and high aspect ratio were obtained. The BaTiO3 plates show good uniformity without obvious agglomeration. The average length of these (001) oriented plates is 6.79 μm. EDS results demonstrate the homogeneously distributed Ba, Ti, and O elements in the plates. XRD patterns, Raman spectra, and PLM results indicate that as-obtained BaTiO3 plates possess a tetragonal phase structure at room temperature. Single-crystal-like extinction behavior and good uniformity are supported by PLM observations. The typical a-c domains are demonstrated by PFM, which is the first observation of typical 90° banded domains in such BaTiO3 plates in mesoscale. The width of banded domains (~125 nm) generally satisfies Kittel’s law of bulk single crystal. Moreover, there are regions with extremely weak signals in comparison with the banded 90° domain regions, which may have potential influences on applications used in the surface performance. The results may provide a new understanding of the domain structure of BaTiO3 mesoscale materials and are helpful for promoting related applications.

Author Contributions

Conceptualization, J.Z. (Jian Zhuang), L.W. and W.R.; Methodology, K.Z. and Y.Q.; Formal analysis, Y.Q., J.Z. (Jinyan Zhao), L.W. and Z.W.; Investigation, K.Z., J.Z. (Jinyan Zhao) and L.W.; Resources, Z.W.; Writing—original draft, K.Z.; Writing—review and editing, J.Z. (Jian Zhuang) and W.R.; Funding acquisition, W.R. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the China National Key R&D Program (2021YFB3201800, 2020YFC0122100), Natural Science Foundation of China (Grants No. 12174299), Key Research Project of Shaanxi Province of China (Grant No. 2018ZDXM-GY-150), Xijiang Innovation Team Introduction Program of Zhaoqing (Jiecheng), China Postdoctoral Science Foundation (Grants No. 2019M663697), the “111 Project” of China (B14040), Funding from National Key Laboratory of Science and Technology on Space Microwave (HTKJ2021KL504015).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare that they have no known competing financial interest or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Von Hippel, A. Ferroelectricity, Domain Structure, and Phase Transitions of Barium Titanate. Rev. Mod. Phys. 1950, 22, 221–237. [Google Scholar] [CrossRef]
  2. Acosta, M.; Novak, N.; Rojas, V.; Patel, S.; Vaish, R.; Koruza, J.; Rossetti, G.A.; Rödel, J. BaTiO3-Based Piezoelectrics: Fundamentals, Current Status, and Perspectives. Appl. Phys. Rev. 2017, 4, 041305. [Google Scholar] [CrossRef] [Green Version]
  3. Wendari, T.P.; Arief, S.; Mufti, N.; Blake, G.R.; Baas, J.; Suendo, V.; Prasetyo, A.; Insani, A.; Zulhadjri, Z. Lead-Free Aurivillius Phase Bi2LaNb1.5Mn0.5O9: Structure, Ferroelectric, Magnetic, and Magnetodielectric Effects. Inorg. Chem. 2022, 61, 8644–8652. [Google Scholar] [CrossRef] [PubMed]
  4. Janil Jamil, N.H.; Zainuddin, Z.; Hj Jumali, M.H.; Izzuddin, I.; Nadzir, L. Tetragonal Tungsten Bronze Phase Potential in Increasing the Piezoelectricity of Sol-Gel Synthesized (K0.5Na0.5)1-xLixNbO3 Ceramics. Ceram. Int. 2022, 48, 9324–9329. [Google Scholar] [CrossRef]
  5. Huang, C.; Wong-Ng, W.; Liu, W.F.; Zhang, X.N.; Jiang, Y.; Wu, P.; Tong, B.Y.; Zhao, H.; Wang, S.Y. Major Improvement of Ferroelectric and Optical Properties in Na-Doped Ruddlesden-Popper Layered Hybrid Improper Ferroelectric Compound, Ca3Ti2O7. J. Alloys Compd. 2019, 770, 582–588. [Google Scholar] [CrossRef]
  6. Zhang, J.; Wang, C.; Bowen, C. Piezoelectric Effects and Electromechanical Theories at the Nanoscale. Nanoscale 2014, 6, 13314–13327. [Google Scholar] [CrossRef]
  7. Liu, S.; Wang, L.; Feng, X.; Wang, Z.; Xu, Q.; Bai, S.; Qin, Y.; Wang, Z.L. Ultrasensitive 2D ZnO Piezotronic Transistor Array for High Resolution Tactile Imaging. Adv. Mater. 2017, 29, 1606346. [Google Scholar] [CrossRef] [PubMed]
  8. Qian, W.; Zhao, K.; Zhang, D.; Bowen, C.R.; Wang, Y.; Yang, Y. Piezoelectric Material-Polymer Composite Porous Foam for Efficient Dye Degradation via the Piezo-Catalytic Effect. ACS Appl. Mater. Interfaces 2019, 11, 27862–27869. [Google Scholar] [CrossRef]
  9. Choi, W.; Choudhary, N.; Han, G.H.; Park, J.; Akinwande, D.; Lee, Y.H. Recent Development of Two-Dimensional Transition Metal Dichalcogenides and Their Applications. Mater. Today 2017, 20, 116–130. [Google Scholar] [CrossRef]
  10. Messing, G.L.; Poterala, S.; Chang, Y.; Frueh, T.; Kupp, E.R.; Watson, B.H.; Walton, R.L.; Brova, M.J.; Hofer, A.K.; Bermejo, R.; et al. Texture-Engineered Ceramics - Property Enhancements through Crystallographic Tailoring. J. Mater. Res. 2017, 32, 3219–3241. [Google Scholar] [CrossRef] [Green Version]
  11. Moriana, A.D.; Zhang, S. Lead-Free Textured Piezoceramics Using Tape Casting: A Review. J. Mater. 2018, 4, 277–303. [Google Scholar] [CrossRef]
  12. Kalhori, H.; Amaechi, I.C.; Youssef, A.H.; Ruediger, A.; Pignolet, A. Catalytic Activity of BaTiO3 Nanoparticles for Wastewater Treatment: Piezo- or Sono-Driven? ACS Appl. Nano Mater. 2022, 6, 1686–1695. [Google Scholar] [CrossRef]
  13. Tang, Q.; Wu, J.; Kim, D.; Franco, C.; Terzopoulou, A.; Veciana, A.; Puigmartí-Luis, J.; Chen, X.Z.; Nelson, B.J.; Pané, S. Enhanced Piezocatalytic Performance of BaTiO3 Nanosheets with Highly Exposed {001} Facets. Adv. Funct. Mater. 2022, 32, 2–9. [Google Scholar] [CrossRef]
  14. Remeika, J.P. A Method for Growing Barium Titanate Single Crystals. J. Am. Chem. Soc. 1954, 76, 940–941. [Google Scholar] [CrossRef]
  15. Merz, W.J. Double Hysteresis Loop of BaTiO3 at the Curie Point. Phys. Rev. 1953, 91, 513–517. [Google Scholar] [CrossRef]
  16. Liu, D.; Yan, Y.; Zhou, H. Synthesis of Micron-Scale Platelet BaTiO3. J. Am. Ceram. Soc. 2007, 90, 1323–1326. [Google Scholar] [CrossRef]
  17. Poterala, S.F.; Chang, Y.; Clark, T.; Meyer, R.J.; Messinge, G.L. Mechanistic Interpretation of the Aurivillius to Perovskite Topochemical Microcrystal Conversion Process. Chem. Mater. 2010, 22, 2061–2068. [Google Scholar] [CrossRef]
  18. Hao, M.; Fan, G.; Lou, Y.; Wen, Y.; Deng, H.; Wang, F.; Lv, W.; Yuchi, M.; Ding, M. Regular and Uniform-Shaped BaTiO3 Microplates Prepared Using a Modified Precursor. Ceram. Int. 2019, 45, 2338–2344. [Google Scholar] [CrossRef]
  19. Kržmanc, M.M.; Jančar, B.; Uršič, H.; Tramšek, M.; Suvorov, D. Tailoring the Shape, Size, Crystal Structure, and Preferential Growth Orientation of BaTiO3 Plates Synthesized through a Topochemical Conversion Process. Cryst. Growth Des. 2017, 17, 3210–3220. [Google Scholar] [CrossRef]
  20. Su, S.; Zuo, R.; Lv, D.; Fu, J. Synthesis and Characterization of (001) Oriented BaTiO3 Platelets through a Topochemical Conversion. Powder Technol. 2012, 217, 11–15. [Google Scholar] [CrossRef]
  21. Aurivillius, B. Mixed bismuth oxides with layer lattices: I. The structure type of CaNb2Bi2O9. Ark. For. Kemi. 1949, 1, 463–480. [Google Scholar]
  22. Pinczuk, A.; Taylor, W.; Burstein, E.; Lefkowitz, I. The Raman Spectrum of BaTiO3. Solid State Commun. 1967, 5, 429–433. [Google Scholar] [CrossRef]
  23. Huang, L.; Chen, Z.; Wilson, J.D.; Banerjee, S.; Robinson, R.D.; Herman, I.P.; Laibowitz, R.; O’Brien, S. Barium Titanate Nanocrystals and Nanocrystal Thin Films: Synthesis, Ferroelectricity, and Dielectric Properties. J. Appl. Phys. 2006, 100, 034316. [Google Scholar] [CrossRef]
  24. Kharat, S.P.; Gaikwad, S.K.; Nalam, P.G.; Kambale, R.C.; James, A.R.; Kolekar, Y.D.; Ramana, C.V. Effect of Crystal Structure and Phase on the Dielectric, Ferroelectric, and Piezoelectric Properties of Ca2+- and Zr4+-Substituted Barium Titanate. Cryst. Growth Des. 2022, 22, 5571–5581. [Google Scholar] [CrossRef]
  25. Xu, K.; Zhu, G.; Xu, H.; Zhao, Y.; Jiang, K.; Zhang, X.; Yin, H.; Shangguan, M.; Wan, L.; Huang, T. The Colossal Permittivity Effect on BaTiO3 Induced by Different Sinter Atmosphere. Appl. Phys. A 2022, 128, 1044. [Google Scholar] [CrossRef]
  26. Pugachev, A.M.; Zaytseva, I.V.; Surovtsev, N.V.; Krylov, A.S. Anharmonicity and Local Noncentrosymmetric Regions in BaTiO3 Pressed Powder Studied by the Raman Line Temperature Dependence. Ceram. Int. 2020, 46, 22619–22623. [Google Scholar] [CrossRef]
  27. Tagantsev, A.K.; Cross, L.E.; Fousek, J. Domains in Ferroic Crystals and Thin Films; Springer: New York, NY, USA, 2010. [Google Scholar]
  28. Huan, Y.; Wang, X.; Fang, J.; Li, L. 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. [Google Scholar] [CrossRef]
  29. McQuaid, R.G.P.; McGilly, L.J.; Sharma, P.; Gruverman, A.; Gregg, J.M. Mesoscale Flux-Closure Domain Formation in Single-Crystal BaTiO3. Nat. Commun. 2011, 2, 404. [Google Scholar] [CrossRef] [Green Version]
  30. Howell, J.A.; Vaudin, M.D.; Cook, R.F. Orientation, Stress, and Strain in an (001) Barium Titanate Single Crystal with 90° Lamellar Domains Determined Using Electron Backscatter Diffraction. J. Mater. Sci. 2014, 49, 2213–2224. [Google Scholar] [CrossRef]
  31. McGilly, L.; Byrne, D.; Harnagea, C.; Schilling, A.; Gregg, J.M. Imaging Domains in BaTiO3 Single Crystal Nanostructures: Comparing Information from Transmission Electron Microscopy and Piezo-Force Microscopy. J. Mater. Sci. 2009, 44, 5197–5204. [Google Scholar] [CrossRef]
  32. Kittel, C. Theory of the Structure of Ferromagnetic Domains in Films and Small Particles. Phys. Rev. 1946, 70, 965–971. [Google Scholar] [CrossRef]
  33. Kittel, C. Physical theory of ferromagnetic domains. Rev. Mod. Phys. 1949, 21, 541. [Google Scholar] [CrossRef]
  34. Schilling, A.; Adams, T.B.; Bowman, R.M.; Gregg, J.M.; Catalan, G.; Scott, J.F. Scaling of Domain Periodicity with Thickness Measured in BaTiO3 Single Crystal Lamellae and Comparison with Other Ferroics. Phys. Rev. B-Condens. Matter Mater. Phys. 2006, 74, 1–6. [Google Scholar] [CrossRef]
Figure 1. FE-SEM images of (ac) Bi4Ti3O12 precursor, (df) BaBi4Ti4O15 precursor, and (gi) BaTiO3 mesoscale particles synthesized at different temperatures.
Figure 1. FE-SEM images of (ac) Bi4Ti3O12 precursor, (df) BaBi4Ti4O15 precursor, and (gi) BaTiO3 mesoscale particles synthesized at different temperatures.
Crystals 13 00538 g001
Figure 2. Length distributions of (ac) Bi4Ti3O12 precursor, (df) BaBi4Ti4O15 precursor, and (gi) BaTiO3 mesoscale particles synthesized at different temperatures.
Figure 2. Length distributions of (ac) Bi4Ti3O12 precursor, (df) BaBi4Ti4O15 precursor, and (gi) BaTiO3 mesoscale particles synthesized at different temperatures.
Crystals 13 00538 g002
Figure 3. (a) XRD patterns and (b) the Raman spectra of BaTiO3 plates synthesized at 950 °C. The inset in (b) indicates the measuring region in a single template with the light spot marked by a green cross.
Figure 3. (a) XRD patterns and (b) the Raman spectra of BaTiO3 plates synthesized at 950 °C. The inset in (b) indicates the measuring region in a single template with the light spot marked by a green cross.
Crystals 13 00538 g003
Figure 4. The EDS mapping results of the BaTiO3 plates synthesized at 950 °C. (a) The FE-SEM image of the scanning area of EDS mapping includes a single BaTiO3 plate. The distribution of (b) Ba element, (c) Ti element, and (d) O element in the scanning area.
Figure 4. The EDS mapping results of the BaTiO3 plates synthesized at 950 °C. (a) The FE-SEM image of the scanning area of EDS mapping includes a single BaTiO3 plate. The distribution of (b) Ba element, (c) Ti element, and (d) O element in the scanning area.
Crystals 13 00538 g004
Figure 5. (a) The photograph of BaTiO3 plates. The PLM images of BaTiO3 micron plates with the crossed polarizers at (b) 0°, (c) 45°, and (d) 90°.
Figure 5. (a) The photograph of BaTiO3 plates. The PLM images of BaTiO3 micron plates with the crossed polarizers at (b) 0°, (c) 45°, and (d) 90°.
Crystals 13 00538 g005
Figure 6. Domain structures of BaTiO3 plate. (a) Out-of-plane PFM image of the whole plate. (b) 3D schematic domain structure. Out-of-plane (c) amplitude and (d) phase image, in-plane (e) amplitude, and (f) phase image of a 6 μm × 6 μm square region.
Figure 6. Domain structures of BaTiO3 plate. (a) Out-of-plane PFM image of the whole plate. (b) 3D schematic domain structure. Out-of-plane (c) amplitude and (d) phase image, in-plane (e) amplitude, and (f) phase image of a 6 μm × 6 μm square region.
Crystals 13 00538 g006
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

Zheng, K.; Zhuang, J.; Quan, Y.; Zhao, J.; Wang, L.; Wang, Z.; Ren, W. The Synthesis and Domain Structures of Single-Crystal-Like Mesoscale BaTiO3 Plates. Crystals 2023, 13, 538. https://doi.org/10.3390/cryst13030538

AMA Style

Zheng K, Zhuang J, Quan Y, Zhao J, Wang L, Wang Z, Ren W. The Synthesis and Domain Structures of Single-Crystal-Like Mesoscale BaTiO3 Plates. Crystals. 2023; 13(3):538. https://doi.org/10.3390/cryst13030538

Chicago/Turabian Style

Zheng, Kun, Jian Zhuang, Yi Quan, Jinyan Zhao, Lingyan Wang, Zhe Wang, and Wei Ren. 2023. "The Synthesis and Domain Structures of Single-Crystal-Like Mesoscale BaTiO3 Plates" Crystals 13, no. 3: 538. https://doi.org/10.3390/cryst13030538

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