Next Article in Journal
On the Applicability of Stereological Methods for the Modelling of a Local Plastic Deformation in Grained Structure: Mathematical Principles
Next Article in Special Issue
Structural and Electric Properties of MnO2-Doped KNN-LT Lead-Free Piezoelectric Ceramics
Previous Article in Journal
Enhanced Performance of Concrete Composites Comprising Waste Metalised Polypropylene Fibres Exposed to Aggressive Environments
Previous Article in Special Issue
Piezoelectric Resonators Excited by Lateral Electric Fields Based on a LiTaO3 Single Crystal
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Spontaneous Growth of CaBi4Ti4O15 Piezoelectric Crystal Using Mixed Flux Agents

1
State Key Laboratory of Crystal Materials and Institute of Crystal Materials, Shandong University, Jinan 250100, China
2
Center for Optics Research and Engineering, Shandong University, Qingdao 266237, China
*
Authors to whom correspondence should be addressed.
Crystals 2020, 10(8), 698; https://doi.org/10.3390/cryst10080698
Submission received: 30 June 2020 / Revised: 9 August 2020 / Accepted: 10 August 2020 / Published: 12 August 2020

Abstract

:
The bismuth layer-structured ferroelectrics (BLSFs) materials have potential for high-temperature piezoelectric applications. Among these piezoelectric materials, the CaBi4Ti4O15 (CBT) piezoelectric ceramic with a high decomposition temperature of about 1250 °C attracts a lot of attention. Achieving a CBT single crystal is a significant way to improve its piezoelectric properties. For this purpose, the flux system for growing CBT crystal was explored in this study. The optimum flux composition ratio was found to be PbO:B2O3:CBT = 3:3:1 in mol%, where the PbO–B2O3 mixtures were used as a flux system. Millimeter size flake-shaped CBT crystals were obtained using the spontaneous growth process for the first time. The relationship between the crystal structure and flake growth habit was analyzed. In addition, the bandgap was evaluated by the combination of transmittance spectrum and first-principle calculations. Besides, the piezoelectric property was predicted from the perspective of polyhedral distortion, which indicated the potential of CBT crystal for piezoelectric applications.

Graphical Abstract

1. Introduction

Since the discovery of the piezoelectric effect, piezoelectric materials have been widely explored as sensors, transducers, actuators, energy harvesting devices, etc. Recently, piezoelectric materials that can maintain the electromechanical and piezoelectric properties at high temperatures and even harsh environments are needed, especially in aerospace industries, oil exploitation, and power plant fields [1,2,3]. The ever-increasing demands on operational temperature have stimulated a great deal of research effort on exploring piezoelectric materials with a high curie temperature (TC) or high melting point, prior to which the electromechanical and piezoelectric properties should be maintained.
To date, several kinds of piezoelectric crystals with high melting points have been studied as the candidates for high-temperature piezoelectric applications [4,5,6]. For example, the modified langasite type crystal Ca3Ta(Ga1−xAlx)3Si2O14 (CTGAS) with a melting point above 1400 °C was determined to possess a relatively high piezoelectric coefficient (d11 = 4.6 pC/N) and electromechanical coupling factor (k12 = 14.4%) [4]. The low-cost calcium aluminate silicate Ca2Al2SiO7 (CAS) crystal with moderate piezoelectric activity was explored for high-temperature pressure sensing [5]. Furthermore, the RECa4O(BO3)3 (ReCOB) crystals with good temperature stability and high electrical resistivity were studied for high-temperature vibration sensing applications [6]. Through the success fabrication of different prototypes of piezoelectric sensors, the sensitivity is expected to be further enhanced, which encourages the exploration of new piezoelectric crystals with strong piezoelectric response. Bismuth layer-structured ferroelectrics (BLSFs), also known as Aurivillius-type ferroelectrics, have attracted much attention due to their strong piezoelectric response and ultrahigh Curie temperature TC (600–900 °C). Among this type of piezoelectric material, the orthorhombic CaBi4Ti4O15 (CBT) compound with a point group of 2 mm is believed to be a promising candidate for high-temperature piezoelectric applications, due to the merits of low electric conductivity, low dielectric loss, relatively large piezoelectric coefficient, and ease of fabrication by using the conventional solid-state reaction method [7,8,9]. However, the practical applications of CBT piezoelectric material at elevated temperatures are still restricted. The piezoelectric activity (d33 = ~7 pC/N) and stability need further improvement [10,11]. Many efforts have been made to improve the electromechanical and piezoelectric properties—for example, preparing the CBT-textured ceramics or conducting ion doping, etc. The piezoelectric coefficient d33 of CBT ceramics prepared by the reactive templated grain growth (RTGG) method with tape casting is reported to be as high as 45 pC/N [12]. However, the achieved improvement in piezoelectric activity for CBT ceramics has the problems of aging easily and grain boundaries. Single crystallization might be a good means to solve the above drawbacks and enhance the piezoelectric properties of CBT compounds. This strategy has been successfully performed on other piezoelectric ceramics. For example, (K0.5Na0.5)NbO3 (KNN) ceramics possessed a low piezoelectric coefficient d33 < 100 pC/N, while the piezoelectric coefficient of the KNN crystal reached about 150 pC/N [13,14].
It is of great significance to perform single crystallization for CBT ceramic and evaluate the electro-elastic properties for potential high-temperature piezoelectric applications. Since the CBT compound is incongruent and would decompose at temperature near 1250 °C, the flux method can be attempted to perform the single crystal growth. The flux method is a convenient, universal, and simple means to cultivate single crystals, and it has been widely used to grow various crystals with incongruent melting points, for example, KNN [14] and KBe2BO3F2 (KBBF) crystals [15], etc. Although there have been extensive studies on the flux growth of functional crystal materials, the investigations on the single crystal growth of a CBT compound are very scare [16], and the search for a suitable flux system for growing CBT single crystal is still on the way.
For this purpose, the flux systems suitable for CBT crystal growth were investigated. The CBT single crystals with dimensions up to millimeter size were grown for the first time using the PbO–B2O3 mixed flux agents. In addition, the relationship between crystal growth morphology and crystal structure was analyzed. The transmittance spectrum and the theoretical calculations of the CBT compound were reported, and the crystal polyhedral distortion and net dipole moment were discussed by theoretical calculation.

2. Materials and Methods

2.1. Polycrystalline Preparation

Prior to single crystal growth, the CaBi4Ti4O15 polycrystalline compound was synthesized by a solid-state reaction method. High-purity CaCO3 (4N, Alfa Aesar, Haverhill, MA, USA), Bi2O3 (4N, Aladdin, Shanghai, China), and TiO2 (4N, Alfa Aesar, Haverhill, MA, USA) powders were selected as the starting raw materials, which were weighed according to the stoichiometric ratio. In order to compensate the evaporation of Bi2O3 during the solid-state reaction process, an excess amount of Bi2O3 (1 mol%) was added to the raw materials. For ensuing the synthesis of the CBT polycrystalline phase, we used two-step sintering. Firstly, all the raw materials were fully mixed, ground, and sintered at 800 °C for 10 h to decompose the CaCO3 completely; then, they were ground again, pressed into cylindrical blocks, and sintered at 1000 °C for 20 h.

2.2. Flux Selection and Growth of CBT Crystal

According to the principle of selecting the flux agent, we studied the fluxing behaviors of two kinds of flux systems for crystallizing CBT crystal, i.e., Bi2O3 and PbO–B2O3 flux systems. The Bi2O3 and PbO–B2O3 flux systems have been successfully applied to the growth of other single crystals [16,17]. To avoid the introduction of impurities, we first consider the self-melting system. Hence, Bi2O3 was selected and used as the flux agent. The prepared CBT polycrystalline powders with the extra added Bi2O3 flux agent were mixed in different mole proportions (Bi2O3:CBT = 3:1, 4:1 and 5:1), heated, and melt at 1000–1100 °C for 20 h and then slowly cooled down to room temperature. In addition to the Bi2O3 flux agent, PbO–B2O3 was chosen as the mixed flux system for crystallizing CBT crystal. PbO and B2O3 have their own merits for crystal growth using the flux method; reports indicate that PbO is an effective solvent used for many oxides and complex perovskites [18,19,20,21,22,23,24]. The B2O3 can reduce the melting temperature, increase the solubility of the melt, and create an optimum degree of complex formation and optimum viscosity [25]. Besides, because of its low melting point and high viscosity, B2O3 can also form a sealing layer to prevent the evaporation of PbO in the molten state at elevated temperatures. Based on the above considerations, the PbO–B2O3 mixed flux system was attempted for CBT single crystal growth. Firstly, the prepared CBT polycrystalline powders were mixed with the PbO–B2O3 flux systems in different mole proportions (PbO:B2O3:CBT = 2:2:1, 3:3:1, 4:4:1, 5:1:1, and 5:3:1) to perform the comparative experiments. Secondly, the mixed compounds were heated and melt at 1000–1100 °C for 20 h; then, they were slowly cooled down to room temperature to find the expected crystal phase.
In this study, the CBT crystal was finally grown by the spontaneous nucleation method using PbO–B2O3 as flux. The CBT polycrystalline powders, high-purity PbO (4N, Aladdin, Shanghai, China) and B2O3 (4N, Alfa Aesar, Haverhill, MA, USA), were weighed in a certain mole proportion (the proportion for CBT and PbO–B2O3 mixed flux was PbO:B2O3:CBT = 3:3:1). The powders were mixed thoroughly, ground, and placed into platinum crucibles in a growth furnace, after which the crucible was slowly heated to 1000 °C and maintained at this temperature for 20 h to make the solution melt completely and mix homogeneously. After that, the crucible was cooled slowly to 500 °C with a very low rate of 2–5 °C/h and then to room temperature at 10 °C/h. When the spontaneous crystallization process was finished, the crucible was boiled with nitric acid to separate the flux from the crystallized CBT crystals.

2.3. Transmission Spectra Measurement and Phase Characterization

The transmission spectra of polycrystalline CBT were measured with a Nicolet IS 10 (Thermo Fisher, Waltham, MA, USA) over the wavelength range of 400 to 4000 nm at room temperature. The crystal phase of the solidified CBT solution obtained by using different flux systems and different proportions was identified by using the Bruker AXS D8 Advance X-ray diffractometer (XRPD, Bruker-axs, Karlsruhe, Germany) at room temperature.

2.4. Theoretical Calculations

The band structure calculations of CBT compound employed the Vienna ab initio Simulation Package (VASP 5.3.5) [26,27] implementation of density functional theory (DFT) in conjunction with the projector-augmented wave (PAW) [28] method. Thereby, the Ca 4s2, Bi 6s26p3, Ti 3d24s2, and O 2s22p4 states were treated as valence electrons. The electronic wave functions were expanded in plane waves using an energy cutoff of 400 eV. The electron exchange and correlation (XC) within the General Gradient Approximation (GGA) of the Perdew, Burke, and Ernzerhof (PBE) [29] functional was used to optimize the configurations, and the force convergence criterion for the structural relaxation was set to 0.01 eV/Å. The electronic properties were performed by the screened hybrid functional of Heyd, Scuseria, and Ernzerhof (HSE06) [30,31]. 4 × 4 × 4 and 2 × 2 × 2 Monkhorst-Pack [32] k-point meshes were employed to sample the Brillouin zones for PBE and HSE06 calculations, respectively.

3. Results and Discussion

3.1. XRPD Analysis and Crystallization

The polycrystalline CBT compounds were synthesized in a stoichiometric ratio with an excess amount of Bi2O3 (1 mol%) added for compensating the evaporation of Bi2O3 during the solid-state reaction process. There were individual impurity peaks (TiO2 component) observed in the compounds obtained by sintering the stoichiometric raw materials using the solid-state reaction method, as shown in Figure 1a. In addition, experimental results proved that the CBT crystalline phase was difficult to obtain solely using Bi2O3 or PbO as flux agents. When using Bi2O3 as the flux agent (Bi2O3:CBT = 3:1, 4:1 and 5:1), the main compound was Bi12TiO20, and no CBT phase was detected using XRPD (Figure 1b). Similarly, when using PbO as the flux agent (PbO:CBT = 2:1, 3:1 and 4:1), the impurity phase is mainly PbxBiyTizO(2x + 3y + 4z)/2 (x, y, z ≥ 0) compounds (Figure 1c). In contrast, when using the mixed PbO–B2O3 flux system and selecting the ratio of PbO:B2O3 to be 1:1, the CBT single crystal phase could be achieved. However, when the ratio deviated from 1:1, the impurity phases such as the Pb5B8O17 and Bi12TiO20 might be formed, as presented in Figure 1d.
Considering that the PbBi4Ti4O15 (PBT) compound has the same symmetry with CBT crystal, further XRPD analysis was performed to identity the crystalline CBT phase. Figure 2a gives the XRPD patterns of crystalline CBT with different flux ratios and compared with the PBT phase. It is observed that the diffraction angle of the main peak at approximately 30.54° for CBT is higher than that of PBT (ICSD No.173544). However, the main peaks in the patterns of crystalline CBT with different flux ratios (PbO:B2O3:CBT = 2:2:1, 3:3:1 and 4:4:1) are kept the same with standard CBT (ICSD No. 99500), supporting that the obtained crystal phase is CBT rather than PBT. Experimental results indicate that when the flux ratio is 2:2:1, the mixed compounds are difficult to solute, while when the ratio is 4:4:1, the solubility is relatively low. Therefore, the optimum flux ratio was obtained to be PbO:B2O3:CBT = 3:3:1. In order to evaluate the purity of crystalline CBT compounds obtained by using the mixed PbO–B2O3 flux agents (PbO:B2O3:CBT = 3:3:1), the refined crystal structure by the Rietveld method for the powder X-ray diffraction was performed. Results are shown in Figure 2b. It is indicated that the differences between the pristine data and the refined data are very small, showing a good agreement with the measurements.
The CBT single crystals with millimeter dimensions were successfully grown by using the spontaneous nucleation process using the mixed PbO–B2O3 flux agents (PbO:B2O3:CBT = 3:3:1). The crystalline state in the platinum crucible and the photographs of as-grown flake-shaped CBT crystals are shown in Figure 3a,b, respectively.

3.2. Structure–Morphology Relationships

The crystal structures of the four-layer ferroelectric Aurivillius phases ABi4Ti4O15 (A = Ca, Sr, Ba, and Pb) were reported by B.J. Kennedy et al. [33], but they lack descriptions in detail about the CBT crystal with a bismuth-layered structure. The relationship between the crystal structure and morphology was analyzed based on the refined structure data by the Rietveld method. The results are shown in Figure 4. The CBT crystal structure is composed of Bi2O2 layers and TiO6 octahedrons arranged alternately along the crystallographic c-axis, where 0.6Bi(2)/0.4Ca(1) and 0.7Bi(3)/0.3Ca(3) atoms are occupied in the intervals among TiO6 octahedrons. The flake-shaped morphology is closely related to the bismuth layer structure of CBT crystal. It is believed that the flake-shaped morphology is associated with the Bi2O2 layer in the CBT crystal. Generally speaking, the bonding strength decreases with the increase of bond length, and large bond lengths are easy to break. It is found that the bond distance of the Bi(1)–O is about 2.55 Å and 2.69 Å, which is much longer than that of the Ti–O bond (1.77 Å and 1.89 Å), as presented at Figure 4b. Therefore, the CBT crystal is relatively difficult to grow along the crystallographic c-axis due to the weak chemical bonding, but it is relatively easy to grow along the a- and b-axes. In addition, it is found that the axis ratio c/a for CBT crystal reached 7.5065, exhibiting strong structural anisotropy. There are the reasons why the CBT crystal grown along the crystallographic c-axis shows a flake growth habit.

3.3. Piezoelectric Activity Evaluation

CaBi4Ti4O15 crystal belongs to the 2 mm point group; thus, it has a non-centrosymmetric structure, which determines that the CBT crystal has piezoelectric properties. In order to explore its potential applications in piezoelectric fields, preliminary evaluations were conducted.

3.3.1. Transmission Spectra

For high-temperature piezoelectric application, the electrical conductivity is an important factor that influences the dielectric loss and temperature stability of the electro-elastic properties. Usually, the electrical conductivity at high temperatures is associated with the bandgap. In this study, the transmittance for CBT polycrystalline sample was measured at room temperature, and the band structure was calculated. As shown in Figure 5, the transmittance from 910 to 4000 nm is about 90% and gradually decreases below the wavelength of 910 nm and reaches zero at about 400 nm, indicating that the transmission cut-off edge of the CBT is about 400 nm, which corresponds to the bandgap of 3.11 eV and is very close to calculated bandgap 3.17 eV. These results are basically consistent with the value reported by Tanwar et al. (3.36 eV) [34].

3.3.2. Dipole Moment Calculation

The piezoelectric activity is closely related to the polyhedron distortion and net dipole moment. As for the polyhedron distortion in CBT crystal, there are several types of polyhedron distortions (Figure 6a), among which the distortion ∆d for Bi(1)O6 octahedron was calculated by using the following formula [35]:
Δ d = | | ( B i O 1 ) | | ( B i O 4 ) | | | cos θ 1 | + | | ( B i O 2 ) | | ( B i O 5 ) | | | cos θ 2 | + | | ( B i O 3 ) | | ( B i O 6 ) | | | cos θ 3 |
where the O1–O6 are coordination oxygen atoms of the Bi(1)O6 octahedron, and θ1, θ2, and θ3 are bond angles. The distortion ∆d for the Bi(1)O6 octahedron was found to be 1.37 Å, which can be identified as strong distortions according to the fourth category defined by Halasyamani [35]. Generally, the larger the polyhedron distortion, the stronger the piezoelectric activity of the crystals. Besides the polyhedron distortion, the net diploe moments in CBT crystal were studied, according to the bond-valence theory [36,37,38,39,40].
Table 1 summarizes the dipole moments of each polyhedron in CBT crystal. It was found that the magnitude of the net dipole moment in the unit cell of the CBT crystal was about 37 Debye, which is comparable with the α-BiB3O6 single crystal (35 Debye), of which the piezoelectric charge coefficient d22 was determined to be in the order of 40 pC/N [41]. Additionally, the directions of the dipole moment for the polyhedral and unit cell of the CBT crystal are identified and shown in Figure 6b, where the net dipole moment direction of the CBT unit cell was negative along the crystallographic c-axis. The calculated results indicated that the Bi(1)O6 octahedron showed a larger distortion and contributed to the piezoelectric activity of the CBT crystal. Hence, it is predicted that the CBT single crystal has excellent piezoelectric performances.

4. Conclusions

In summary, the flake-shaped CBT crystals in the millimeter range were grown by the flux method using the mixed PbO–B2O3 agents for the first time. The optimum flux composition mole ratio was found to be PbO:B2O3:CBT = 3:3:1. The flake-shaped growth habit of the CBT crystal was analyzed from the perspective of a microcrystal structure. The structural distortions and the net dipole moment of the CBT crystal were found to be 1.37 Å (Bi(1)O6 octahedron) and 37 Debye, respectively. The Bi(1)O6 octahedron showed the larger distortion and believed to contribute to the piezoelectric activity of the CBT crystal. In addition, the transmittance spectrum of the CBT crystal was measured and found to show a broad transmission range, where the bandgap was found to be near 3.11 eV. However, the main problem encountered in CBT growth is the observed flake growth habit. In the following studies, the large-size single crystal growth by using the top-seeded method with mixed PbO–B2O3 fluxes will be performed.

Author Contributions

Conceptualization, M.F. and F.Y.; methodology, M.F., L.L., Y.L. and X.C.; validation, F.Y., X.Z.; investigation, M.F. and G.W.; resources, F.Y. and X.Z.; writing—original draft preparation, M.F.; writing—review and editing, F.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by The Key Research and Development Program of Shandong Province (Nos.2017CXGC0413 and 2019JZZY010313), and the National Natural Science Foundation of China (No. 51872165).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, S.J.; Yu, F.P. Piezoelectric materials for high temperature sensors. J. Am. Ceram. Soc. 2011, 94, 3153–3170. [Google Scholar] [CrossRef]
  2. Jiang, X.N.; Kim, K.; Zhang, S.J.; Johnson, J.; Salazar, G. High-temperature piezoelectric sensing. Sensors 2014, 14, 144–169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Wu, J.G.; Gao, X.Y.; Chen, J.G.; Wang, C.M.; Zhang, S.J.; Dong, S.X. Review of high temperature piezoelectric materials, devices, and applications. Acta Phys. Sin. 2018, 67, 207701. [Google Scholar]
  4. Yokota, Y.; Ohashi, Y.; Kudo, T.; Kochurikhin, V.V.; Medvedev, A.; Kurosawa, S.; Kamada, K.; Yoshikawa, A. Effects of Al Substitution for Ca3Ta(Ga1−xAlx)3Si2O14 Piezoelectric Single Crystals. J. Cryst. Growth 2017, 468, 321–325. [Google Scholar] [CrossRef]
  5. Hagiwara, M.; Noguchi, H.; Hoshina, T.; Takeda, H.; Fujihara, S.; Kodama, N.; Tsurumi, T. Growth and Characterization of Ca2Al2SiO7 Piezoelectric Single Crystals for High-Temperature Sensor Applications. Jpn. J. Appl. Phys. 2013, 52, 09KD03. [Google Scholar] [CrossRef]
  6. Zhang, S.J.; Fei, Y.T.; Frantz, E.; Snyder, D.W.; Chai, B.H.T.; Shrout, T.R. High-Temperature Piezoelectric Single Crystal ReCa4O(BO3)3 for Sensor Applications. IEEE Trans. Ultrason. Ferroelectr. Freq. Control 2008, 55, 2703–2708. [Google Scholar] [CrossRef]
  7. Zheng, L.Y.; Li, G.R.; Zhang, W.Z.; Yin, Q.R. The Structure and Properties of Bi-Layered Piezoelectric Ceramics Bi4(Ca, Sr)Ti4O15. Jpn. J. Appl. Phys. 2002, 41, L1471–L1473. [Google Scholar] [CrossRef]
  8. Yan, H.X.; Li, C.E.; Zhou, J.G.; Zhu, W.M.; He, L.X.; Song, Y.X.; Yu, Y.H. Effects of A-Site (NaCe) Substitution with Na-Deficiency on Structures and Properties of CaBi4Ti4O15-Based High-Curie-Temperature Ceramics. Jpn. J. Appl. Phys. 2001, 40, 6501–6505. [Google Scholar] [CrossRef]
  9. Subbarao, E.C. A Family of Ferroelectric Bismuth Compounds. J. Phys. Chem. Solids 1962, 23, 665–676. [Google Scholar] [CrossRef]
  10. Khokhar, A.; Mahesh, M.L.V.; James, A.R.; Goyal, P.K.; Sreenivas, K. Sintering Characteristics and Electrical Properties of BaBi4Ti4O15 Ferroelectric Ceramics. J. Alloy. Compd. 2013, 581, 150–159. [Google Scholar] [CrossRef]
  11. Zhou, Z.Y.; Li, Y.C.; Hui, S.P.; Dong, X.L. Effect of Tungsten Doping in Bismuth-Layered Na0.5Bi2.5Nb2O9 High Temperature Piezoceramics. Appl. Phys. Lett. 2014, 104, 012904. [Google Scholar] [CrossRef]
  12. Takeuchi, T.; Tani, T.; Saito, Y. Piezoelectric Properties of Bismuth Layer-Structured Ferroelectric Ceramics with a Preferred Orientation Processed by the Reactive Templated Grain Growth Method. Jpn. J. Appl. Phys. 1999, 38, 5553–5556. [Google Scholar] [CrossRef]
  13. Jaeger, R.E.; Egerton, L. Hot Pressing of Potassium-Sodium Niobates. J. Am. Ceram. Soc. 1962, 45, 209–213. [Google Scholar] [CrossRef]
  14. Lin, D.B.; Li, Z.R.; Zhang, S.J.; Xu, Z.; Yao, X. Dielectric/Piezoelectric Properties and Temperature Dependence of Domain Structure Evolution in Lead Free (K0.5Na0.5)NbO3 single crystal. Solid State Commun. 2009, 149, 1646–1649. [Google Scholar] [CrossRef]
  15. Tang, D.Y.; Xia, Y.N.; Wu, B.C.; Chen, C.T. Growth of A New UV Nonlinear Optical Crystal: KBe2(BO3)F2. J. Cryst. Growth 2001, 222, 125–129. [Google Scholar] [CrossRef]
  16. Tellier, T.; Boullay, P.; Manier, M.; Mercurio, D. A Comparative Study of the Aurivillius Phase Ferroelectrics CaBi4Ti4O15 and BaBi4Ti4O15. J. Solid State Chem. 2004, 177, 1829–1837. [Google Scholar] [CrossRef]
  17. Dong, M.; Ye, Z.G. High-Temperature Solution Growth and Characterization of the Piezo-/Ferroelectric (1−x)Pb(Mg1/3Nb2/3)O3-xPbTiO3 [PMNT] single crystals. J. Cryst. Growth 2000, 209, 81–90. [Google Scholar] [CrossRef]
  18. Saitoh, S.; Takeuchi, T.; Kobayashi, T.; Harada, K.; Shimanuki, S.; Yamashita, Y. An Improved Phased Array Ultrasonic Probe Using 0.91Pb(Zn1/3Nb2/3)O3-0.09PbTiO3 Single Crystal. Jpn. J. Appl. Phys. 1999, 38, 3380–3384. [Google Scholar] [CrossRef]
  19. Yamashita, Y.; Shimanuki, S. Synthesis of Lead Scandium Niobate-Lead Titanate Pseudo Binary System Single Crystals. Mater. Res. Bull. 1996, 31, 887–895. [Google Scholar] [CrossRef]
  20. Brunskill, I.H.; Boutellier, R.; Depmeier, W.; Schmid, H. High-Temperature Solution Growth of Pb(Fe0.5Nb0.5)O3 and Pb(Mn0.5Nb0.5)O3 Crystals. J. Cryst. Growth 1982, 56, 541–546. [Google Scholar] [CrossRef]
  21. Brixel, W.; Boutellier, R.; Schmid, H. Flux Growth and Characterization of Single Crystals of the Perovskites Pb2FeTaO6 and Pb2CoWO6. J. Cryst. Growth 1987, 82, 396–404. [Google Scholar] [CrossRef] [Green Version]
  22. Ye, Z.G.; Tissot, P.; Schmid, H. Pseudo-Binary Pb(Mg1/3Nb2/3)O3-PbO Phase Diagram and Crystal Growth of Pb(Mg1/3Nb2/3)O3 [PMN]. Mater. Res. Bull. 1990, 25, 739–748. [Google Scholar] [CrossRef]
  23. Sun, B.N.; Boutellier, R.; Sciau, P.; Burkhardt, E.; Rodriguez, V.; Schmid, H. High Temperature Solution Growth of Perovskite Pb2CoWO6 Single Crystals. J. Cryst. Growth 1991, 112, 71–83. [Google Scholar] [CrossRef]
  24. Ye, Z.G.; Schmid, H. Growth From High Temperature Solution and Characterization of Pb(Fe2/3W1/3)O3 Single Crystals. J. Cryst. Growth 1996, 167, 628–637. [Google Scholar] [CrossRef]
  25. Elwell, D.; Sheel, H.J. Crystal Growth from High Temperature Solution, 3rd ed.; Academic Press: New York, NY, USA, 1975; pp. 94–95. [Google Scholar]
  26. Kresse, G.; Furthmüller, J. Efficiency of ab-initio Total Energy Calculations for Metals and Semiconductors Using a Plane-Wave Basis Set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  27. Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for ab initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  28. Kresse, G.; Joubert, D. From Ultrasoft Pseudopotentials to the Projector Augmented-Wave Method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  29. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [Green Version]
  30. Heyd, J.; Scuseria, G.E.; Ernzerhof, M. Hybrid Functionals Based on a Screened Coulomb Potential. J. Chem. Phys. 2003, 118, 8207–8215. [Google Scholar] [CrossRef] [Green Version]
  31. Krukau, A.V.; Vydrov, O.A.; Izmaylov, A.F.; Scuseria, G.E. Influence of the Exchange Screening Parameter on the Performance of Screened Hybrid Functionals. J. Chem. Phys. 2006, 125, 224106. [Google Scholar] [CrossRef]
  32. Monkhorst, H.J.; Pack, J.D. Special Points for Brillouin-Zone Integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  33. Kennedy, B.J.; Zhou, Q.D.; Ismunandar; Kubota, Y.; Kato, K. Cation Disorder and Phase Transitions in the Four-Layer Ferroelectric Aurivillius Phases ABi4Ti4O15 (A = Ca, Sr, Ba, Pb). J. Solid State Chem. 2008, 181, 1377–1386. [Google Scholar] [CrossRef]
  34. Tanwar, A.; Sreenivas, K.; Gupta, V. Effect of Orthorhombic Distortion on Dielectric and Piezoelectric Properties of CaBi4Ti4O15 Ceramics. J. Appl. Phys. 2009, 105, 084105. [Google Scholar] [CrossRef]
  35. Halasyamani, P.S. Asymmetric Cation Coordination in Oxide Materials: Influence of Lone-Pair Cations on the Intra-Octahedral Distortion in d0 Transition Metals. Chem. Mater. 2004, 16, 3586–3592. [Google Scholar] [CrossRef]
  36. Brese, N.E.; O’Keeffe, M. Bond-Valence Parameters for Solids. Acta Cryst. B 1991, 47, 192–197. [Google Scholar] [CrossRef]
  37. Ok, K.M.; Halasyamani, P.S. Mixed-Metal Tellurites: Synthesis, Structure, and Characterization of Na1.4Nb3Te4.9O18 and NaNb3Te4O16. Inorg. Chem. 2005, 44, 3919–3925. [Google Scholar] [CrossRef]
  38. Maggard, P.A.; Nault, T.S.; Stern, C.L.; Poeppelmeier, K.R. Alignment of Acentric MoO3F33− Anions in a Polar Material: (Ag3MoO3F3)(Ag3MoO4)Cl. J. Solid State Chem. 2003, 175, 27–33. [Google Scholar] [CrossRef]
  39. Izumi, H.K.; Kirsch, J.E.; Stern, C.L.; Poeppelmeier, K.R. Examining the Out-of-Center Distortion in the [NbOF5]2− Anion. Inorg. Chem. 2005, 44, 884–895. [Google Scholar] [CrossRef]
  40. Kim, J.H.; Halasyamani, P.S. A Rare Multi-Coordinate Tellurite, NH4ATe4O9·2H2O (A = Rb or Cs): The Occurrence of TeO3, TeO4, and TeO5 Polyhedra in the Same Material. J. Solid State Chem. 2008, 181, 2108–2112. [Google Scholar] [CrossRef]
  41. Chen, F.F.; Jiang, C.; Tian, S.W.; Yu, F.P.; Cheng, X.F.; Duan, X.L.; Wang, Z.P.; Zhao, X. Electroelastic Features of Piezoelectric Bi2ZnB2O7 Crystal. Cryst. Growth Des. 2018, 18, 3988–3996. [Google Scholar] [CrossRef]
Figure 1. X-ray diffractometer (XRPD) patterns of the crystalline CaBi4Ti4O15 (CBT) compounds: (a) XRPD pattern of the synthesized polycrystalline powder; (b) XRPD pattern of the polycrystalline using Bi2O3 as flux; (c) XRPD pattern of the polycrystalline using PbO as flux; (d) XRPD pattern of polycrystalline using PbO–B2O3 as flux.
Figure 1. X-ray diffractometer (XRPD) patterns of the crystalline CaBi4Ti4O15 (CBT) compounds: (a) XRPD pattern of the synthesized polycrystalline powder; (b) XRPD pattern of the polycrystalline using Bi2O3 as flux; (c) XRPD pattern of the polycrystalline using PbO as flux; (d) XRPD pattern of polycrystalline using PbO–B2O3 as flux.
Crystals 10 00698 g001
Figure 2. (a) XRPD patterns of crystalline CBT with different flux ratios (PbO:B2O3:CBT = 2:2:1, 3:3:1, and 4:4:1) and compared with the PbBi4Ti4O15 (PBT) phase; (b) XRPD patterns of crystalline CBT compounds obtained after refinement.
Figure 2. (a) XRPD patterns of crystalline CBT with different flux ratios (PbO:B2O3:CBT = 2:2:1, 3:3:1, and 4:4:1) and compared with the PbBi4Ti4O15 (PBT) phase; (b) XRPD patterns of crystalline CBT compounds obtained after refinement.
Crystals 10 00698 g002
Figure 3. (a) Top-view of the as-cooled Pt crucible showing the grown CBT crystals by spontaneous nucleation process; (b) Selected flake-shaped CBT crystals crystallized from mixed flux agents (PbO:B2O3:CBT = 3:3:1).
Figure 3. (a) Top-view of the as-cooled Pt crucible showing the grown CBT crystals by spontaneous nucleation process; (b) Selected flake-shaped CBT crystals crystallized from mixed flux agents (PbO:B2O3:CBT = 3:3:1).
Crystals 10 00698 g003
Figure 4. Schematic crystal structure of the CaBi4Ti4O15 crystal: (a) projected along the [110] direction; (b) projected along the crystallographic a-axis in space (the unit in Å).
Figure 4. Schematic crystal structure of the CaBi4Ti4O15 crystal: (a) projected along the [110] direction; (b) projected along the crystallographic a-axis in space (the unit in Å).
Crystals 10 00698 g004
Figure 5. The transmittance spectrum of the prepared polycrystalline CBT powders measured at room temperature (inset shows the calculated band structure).
Figure 5. The transmittance spectrum of the prepared polycrystalline CBT powders measured at room temperature (inset shows the calculated band structure).
Crystals 10 00698 g005
Figure 6. (a) Structural distortions for MO6 octahedron in the CBT crystal (M represents Bi, Ca, and Ti atoms); (b) Directions of the dipole moment in the polyhedral and unit cell of CBT crystal (the black arrow indicates the net dipole moment).
Figure 6. (a) Structural distortions for MO6 octahedron in the CBT crystal (M represents Bi, Ca, and Ti atoms); (b) Directions of the dipole moment in the polyhedral and unit cell of CBT crystal (the black arrow indicates the net dipole moment).
Crystals 10 00698 g006
Table 1. Dipole moments calculated for CBT single crystal.
Table 1. Dipole moments calculated for CBT single crystal.
CrystalSpecies∆d (Å)Dipole Moments
x (a)y (b)z (c)Debye×10−4 edu·cm/Å3
CBT *Bi(1)O61.37−2.0168−6.7630−14.540616.16541.46
Ca(1)/Bi(2)O80.60−9.9685−0.144709.97333.99
Ca(3)/Bi(3)O7-−8.22393.390410.849114.03470.00
Ti(1)O60.444.34570.0852−2.84555.20174.04
Ti(2)O60.692.3948−0.5329−8.49058.84296.07
Unit cell-00−37.35−37.35−312.78
* The crystallographic data of CBT was derived from the refined structure by the Rietveld method.

Share and Cite

MDPI and ACS Style

Fan, M.; Wu, G.; Yu, F.; Li, L.; Li, Y.; Cheng, X.; Zhao, X. Spontaneous Growth of CaBi4Ti4O15 Piezoelectric Crystal Using Mixed Flux Agents. Crystals 2020, 10, 698. https://doi.org/10.3390/cryst10080698

AMA Style

Fan M, Wu G, Yu F, Li L, Li Y, Cheng X, Zhao X. Spontaneous Growth of CaBi4Ti4O15 Piezoelectric Crystal Using Mixed Flux Agents. Crystals. 2020; 10(8):698. https://doi.org/10.3390/cryst10080698

Chicago/Turabian Style

Fan, Mengdi, Guangda Wu, Fapeng Yu, Lili Li, Yanlu Li, Xiufeng Cheng, and Xian Zhao. 2020. "Spontaneous Growth of CaBi4Ti4O15 Piezoelectric Crystal Using Mixed Flux Agents" Crystals 10, no. 8: 698. https://doi.org/10.3390/cryst10080698

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