Next Article in Journal
Study of Bond Strength of Steel Bars in Basalt Fibre Reinforced High Performance Concrete
Next Article in Special Issue
Effect of Chitosan Electrospun Fiber Mesh as Template on the Crystallization of Calcium Oxalate
Previous Article in Journal / Special Issue
Origin of Structural Change Driven by A-Site Lanthanide Doping in ABO3-Type Perovskite Ferroelectrics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ultrahigh Electromechanical Coupling and Its Thermal Stability in (Na1/2Bi1/2)TiO3-Based Lead-Free Single Crystals

1
Jiangxi Key Laboratory of Advanced Ceramic Materials, National Engineering Research Center for Domestic & Building Ceramics, School of Materials Science and Engineering, Jingdezhen Ceramic Institute, Jingdezhen, Jiangxi 333403, China
2
Institute for Superconducting and Electronic Materials, Australian Institute for Innovative Materials, University of Wollongong, North Wollongong 2500, Australia
3
Artificial Crystal Research Center, Shanghai Institute of Ceramics, University of Chinese Academy of Sciences, Jiading, Shanghai 200050, China
*
Author to whom correspondence should be addressed.
Crystals 2020, 10(6), 435; https://doi.org/10.3390/cryst10060435
Submission received: 7 May 2020 / Revised: 24 May 2020 / Accepted: 27 May 2020 / Published: 29 May 2020

Abstract

:
In this work, we report the ultrahigh electromechanical coupling performance of NBT-6BT-KNN lead-free single crystal at room temperature. The thickness mode electromechanical coupling coefficient (kt) and the 31 mode electromechanical coupling coefficient (k31) reach 69.0% and 45.7%, respectively, which are superior to the PZT-5H lead-based ceramics of kt~60% and k31~39%. In addition, the evolution of the crystal structure and domain morphology is revealed by Raman scattering spectra, a polarizing microscope and piezoelectric force microscopy characterization.

1. Introduction

Piezoelectric materials, which convert mechanical to electrical energy (and vice versa), are crucial in sensors, actuators, transducers, and ultrasonic devices [1,2,3]. Lead zirconium titanate (PZT) based ceramics are currently market-dominating due to their excellent piezoelectric properties [4,5]. However, these PZT family materials contain toxic lead, which is restricted in view of the environment. Great efforts have been made worldwide to seek alternative lead-free piezoelectric materials [6,7,8]. The essential strategy for achieving a high piezoelectric response is to seek lead-free materials displaying a transition region in their composition phase diagrams, known as a morphotropic phase boundary (MPB), where energetically comparable polar states coexist and compete with each other [6,7]. Such critical energy landscapes cause a fragility and instability of the polar state, so that the polarization can be easily rotated and altered by external stress or electric field. Some developments in the search for lead-free piezoelectric materials are BaTiO3-based, (K1/2Na1/2)NbO3-based, and Na1/2Bi1/2TiO3-based systems approached by constructing an MPB [6,9,10]. It was later discovered that an MPB could be also created through external mechanical stress or electrical field, which results in a strong piezoelectric response [11,12]. Electric-field in situ Transmission Electron Microscopy (TEM) has revealed that the overall poling-induced phase transitions in 0.94Na1/2Bi1/2TiO3-0.06BaTiO3 ceramics can be described as R3c/P4bm→P4mm/R3c→R3c [11]. Ge et al. and Daniels et al. proposed that an E-field induced phase transformation from pseudo cubic to T structure is responsible for the large electrically induced strain [13,14]. Recently, Wang et al. [15] reported a ternary lead-free Na1/2Bi1/2TiO3-BaTiO3-(K1/2Na1/2)NbO3 (NBT-BT-KNN) single crystal that exhibits an unprecedented piezoelectric coefficient of d3 ≈ 840 pC N−1. The origin of the high piezoelectricity is attributed to a phase boundary constructed by the poling field between multiple weak-polar R3c/P4bm phases and ferroelectric tetragonal phases with P4mm symmetry.
The electromechanical coupling coefficient is an important parameter which characterizes the coupling between the mechanical energy and electrical energy of piezoelectric materials. However, the electromechanical coupling coefficient of this ternary NBT-BT-KNN single crystal has not yet been presented. In this communication, we report the ultrahigh electromechanical coupling performance of NBT-6BT-KNN single crystal. The thickness mode electromechanical coupling coefficient (kt) and the 31-mode electromechanical coupling coefficient (k31) reach 69.0% and 45.7%, respectively, which are superior to the PZT-5H lead-based ceramics of kt ~ 60% and k31~39%. The temperature-dependent kt and k31 are further studied, which reveals a mechanism for temperature-induced phase transition.

2. Materials and Methods

(0.94−x)(Na1/2Bi1/2)TiO3-0.06BaTiO3-xK1/2Na1/2NbO3 (NBT-6BT-100xKNN, x = 0.01) single crystals were grown by a carefully controlled TSSG technique. The raw materials of K2CO3 (99.99%), Na2CO3 (99.99%), Bi2O3 (99.999%), BaCO3 (99.99%), Nb2O5 (99.99%), and TiO2 (99.99%) were weighed according to the nominal ratio of NBT-6BT-KNN. Then the well-mixed compounds were put into a platinum (Pt) crucible and calcined at 850 °C for 3 h in air to form the NBT-6BT-KNN polycrystalline precursor. Then, the polycrystalline materials were ground and mixed with excess ~20 wt% Na2CO3, K2CO3, and Bi2O3 as a self-flux. Single crystals were grown using a <001> direction seed in a Pt crucible, which was heated to ~1300 °C using a resistance furnace under an air atmosphere. The details of crystal growth can be found elsewhere [15]. The x-ray fluorescence analyzer (XRFA) and x-ray diffraction (XRD) were utilized to determine the real composition and phase structure of the as-grown crystal, respectively.
For the macroscopic electrical characterization, the as-grown single crystals were sliced into wafers with a thickness of ~0.40 mm, perpendicular to their pseudo cubic <100> directions. Based on the IEEE standards, the [001]-oriented crystal planes with a size of 4 × 4 × 0.4 mm3 were cut to obtain thickness mode samples. The 31 mode samples were cut into rectangular parallelepiped shapes along with crystallographic directions of [100]/[010]/[001] with a size of 12.3 × 2.4 × 0.4 mm3. Then, the main faces of the samples were sputtered with gold electrodes. The crystal samples were poled in silicon oil at room temperature for 15 min with a dc electric field of 25 kV cm−1. Dielectric/loss-temperature and impedance-phase angle measurements were conducted using a computer-controlled Novel control CmbH Concept 40 broadband dielectric spectrometer. Raman measurements were conducted at room temperature using a Raman spectrometer (LabRAM HR800, Horiba JobinYvon, Paris, France). The domain structure was observed using an Olympus (Tokyo, Japan) U-CMAD3 polarizing microscope (PLM) equipped with a Linkam THMS600E (Linkam Scientific Instruments Ltd.,Tadworth, Surrey, U.K). The PLM is used in transmission geometry, and the principle of setup in our case was as follows: we defined the “θ” as the angle between the polarizer/analyzer (P/A) pair and the pseudo cubic <100> direction. For the domains of the tetragonal phase, the optical extinction is achieved when the P/A is set along the <100> direction, namely at θ = 0° with respect to the <100> direction. For the domains of the rhombohedral phase, they exhibit optical extinction when the P/A is set along the <110> direction, namely at θ = 45° with respect to the <100> direction. The PFM experiments were carried out using a commercial atomic force microscope MFP-3D (Asylum Research, Goleta, USA). An ac modulation voltage of 7 V (peak to peak) with a frequency of 10 kHz was applied between the conductive tip and the bottom gold electrode.

3. Results and Discussion

Figure 1a,b shows the thickness mode and 31 mode impedance and phase angle of the poled NBT-6BT-KNN crystal along the <001> direction at room temperature, respectively. It is clear from the figure that the phase angle of NBT-6BT-KNN crystal in the two modes is close to 80°, indicating the completely polarized state. The electromechanical coupling coefficients kt and k31 are calculated by the resonance-antiresonance method (as shown in Formula (1), (2)), which are 69.0% and 45.7%, respectively, superior to those of kt ~ 60% and k31-39% of PZT-5H lead-based ceramics [16]. We have also compared our own results in this work to the results from full sets of electromechanical constants of other lead-free ferroelectric (piezoelectric) ceramic/single crystals; details can be seen in Table 1 [16,17,18,19,20,21].
k t = π 2 f r f a tan ( π 2 f a f r f a )
k 31 = 1 1 tan [ ( π / 2 ) ( f a / f r ) ] ( π / 2 ) ( f a / f r )
Figure 1c,d shows the temperature-dependent behavior of kt and k31, respectively. It is clear from the figures that kt and k31 exhibit good temperature stability within this temperature range between −150 and 40 °C, indicating an excellent cryogenic electromechanical coupling property and a promising candidate for advanced electromechanical transducer applications below room temperature. When the temperature reaches above 45 °C, the value of kt and k31 degenerates sharply to zero. We also measured the temperature-dependent dielectric constant and dielectric loss curves of <001> poled oriented NBT-6BT-KNN crystal at different frequencies, as shown in Figure 2. In general, two dielectric anomalies can be observed during the heating process. The temperature corresponding to the maximum dielectric constant is considered to be the phase transition temperature from the tetragonal phase to the tetragonal + cubic two phase region [10], and the dielectric anomaly at low temperature around 150 °C range shows strong relaxation characteristics of dielectric broadening and frequency dispersion, which are closely related to the polar nanoscale (PNRs) transition from the R3c to the tetragonal P4bm phase according to the TEM results of NBT-based ceramics. These PNRs in the R3c or P4bm phase have a strong relaxation and show a weak/non-polar state [22]. It should be noted that a discontinuous variation in dielectric/loss curves can be only observed in poled crystal and the anomaly also appears exactly at 45 °C, defined as depolarization temperature (Td). This is well consistent with the degeneration temperature of kt and k31. The reason for this observation is most likely that the electric field induces the transformation from the weak/non-polar pseudocubic phases (R3c/P4bm) to the polar tetragonal phase (P4mm) and heating causes the tetragonal polar phase to transform back to the non-polar phases [15].
In order to verify this assumption, we observe the domain evolution of <001> oriented NBT-6BT-KNN crystals under different electric fields and temperatures using a polarizing microscope, as shown in Figure 3. At zero electric field (E = 0 kV/cm), NBT-6BT-KNN crystals show complete light extinction, indicating the initial pseudocubic non-polar state. This is consistent with the XRD result of as-grown NBT-6BT-KNN crystal powder (Figure 4). The diffraction pattern shows no profile shape asymmetry or splitting at any reflections, revealing that the sample is an initial cubic or pseudo cubic structure. When the electric field increases to 25 kV/cm, obvious macroscopic ferroelectric domains are present for P/A: 45° (θ = 45°), indicating that the electric field induces the phase transition from a non-polar pseudocubic phase to a ferroelectric polar phase [15,23]. The red dotted line and arrows in Figure 3b show the single domain area and crystal cracks, respectively. When the electric field is removed, the ferroelectric domains are still maintained (Figure 3c), revealing that the polar phase is stable after removing the electric field. In addition, as shown in the inset of Figure 3c, the optical extinction of these domains for P/A: 0° (θ = 0°) indicates that the polar phase belongs to the P4mm symmetry. When the temperature is increased above Td (T = 50 °C), the ferroelectric domain disappears completely, as presented in Figure 3d, which confirms the temperature induced a transition back to the non-polar phases.
Macroscopic Raman spectra contains local information via local phonon signals. Previous studies of NBT-based systems reveal that the changes in these Raman modes can reflect the crystal phase transition sensitively. Figure 5a,b shows the room temperature Raman spectra of unpoled (E = 0 kV/cm) and poled (E = 25 kV/cm) NBT-6BT-KNN crystals, respectively. The Raman spectra have been deconvoluted into six vibrational modes and the assignment of these modes is marked. The bands around 300 cm−1 are assigned to the Ti-O lattice vibration. The bands (around 527 cm−1 and 610 cm−1) between 400 and 700 cm−1 are assigned to the TiO6 octahedron vibration closely related to crystal symmetry [24,25]. After applying an electric field, two changes are clearly noticed. The relative intensity of 527 cm−1 and 610 cm−1 related to TiO6 octahedron vibration changes significantly, and the vibration around the 610 cm−1 band is significantly weakened compared to the mode around 527 cm−1. The behavior in these high frequency modes suggests a reduced degree of disorder [26], which further supports the electric-field-induced phase transition from pseudocubic a state to a ferroelectric tetragonal state. The electric-field-induced TiO6 octahedron dynamics, in turn, affect the Ti-O band. The poled crystal exhibits relatively more symmetry around the 300 cm−1 band, as compared to that of the unpoled crystal, which is also consistent with the reduced degree of disorder. When the poled crystals are heated to a temperature above Td (T = 50 °C), the Raman spectra are then measured and shown in Figure 5c. It is clear that the Raman scattering profile almost completely evolves back to the initial state without poling, indicating that the temperature induced phase transitions from the polar tetragonal state back to the initial nonpolar pseudocubic state. These results are exactly consistent with the domain evolutions in Figure 3.
The sharply degenerated kt and k31 at Td for poled NBT-6BT-KNN crystals can now be well understood by the electric filed/temperature induced phase transition between the non-polar and polar states. When the crystals are poled at 25 kV /cm, the phase transitions occur from the pseudocubic non-polar phase to the tetragonal polar phase. However, when the crystals are heated to a temperature reaching above Td, the ferroelectric polar phases are unstable and transform back to a pseudocubic non-polar phase, accompanied by a sharp degenerated kt and k31 and a discontinuous variation in dielectric/loss curves at exactly Td.
The piezoelectric and electromechanical coupling coefficients are strongly dependent on microstructural changes of nanodomains [27,28]. Piezoelectric force microscopy (PFM) is a very good experimental approach to detect ferroelectric nanoscale domains. In order to trace the origin of the ultrahigh electromechanical coupling coefficients kt and k31 in NBT-6BT-KNN crystals, we used PFM to test the microscopic domain structure of NBT-6BT and NBT-6BT-KNN crystals. Figure 6 shows the PFM images of NBT-6BT-KNN and NBT-6BT crystals along the <001> direction. It can be seen that nano-regions of NBT-6BT-KNN exhibit a more dense and uniform distribution than that of NBT-6BT. In particular, the polar-regions became notably narrower with sizes, on average, of ~50 nm. As reflected in the amplify images (Figure 6c–d), the boundaries between regions became much smoother. These results clearly show that the addition of KNN to NBT-BT refines the size of the polar nanoregions and enhances their self-organization. Since the domain size is proportional to the square root of the domain wall energy, the smaller domain size makes it easily reoriented under external excitations, e.g., applied electric field or mechanical force [29,30]. Thus, the piezoelectric properties significantly improve with a decreasing domain size. It should be noted that these results have been found in BaTiO3 ferroelectric single crystal [31,32,33,34]. Therefore, such refinement of ferroelectric domains in NBT-6BT-KNN crystal would lead to an increased domain-wall density and contribute to its ultrahigh value of electromechanical coupling performance.

4. Conclusions

0.93(Na1/2Bi1/2)TiO3-0.06BaTiO3-0.01K1/2Na1/2NbO3 (NBT-6BT-KNN) single crystals show ultrahigh electromechanical coupling coefficients kt and k31 reaching 69.0% and 45.7% respectively, which are superior to the PZT-5H lead-based ceramics of kt ~ 60% and k31 ~ 39%. kt and k31 vary little and exhibit a good temperature stability within this temperature range from −150 to 40 °C. When the temperature reaches above 45 °C, the value of kt and k31 degenerates sharply to zero, which suggests a temperature induced tetragonal polar to pseudocubic non-polar phase transition verified by dielectric, polarizing microscope and Raman scattering measurements. The refinement of the ferroelectric domains and high domain-wall density was observed in the PFM results, which contributes to the ultrahigh electromechanical coupling of NBT-6BT-KNN crystals.

Author Contributions

For research articles with several authors, a short paragraph specifying their individual contributions must be provided. C.C. gave the onceptualization and wrote original draft preparation; L.Y. did data curation; X.J. (Xingan Jiang), X.H., X.G., N.T., K.S. joined formal analysis; X.J. (Xiangping Jiang), S.Z., H.L. supervised this work. All authors have read and agreed to the published version of the manuscript.

Funding

This work is financially supported by the National Natural Science Foundation of China (51862016, 51762024, 51602135), the Foundation of Jiangxi Provincial Education Department (GJJ190712). The author (Chao Chen) wishes to acknowledge the support from the Jiangxi Voyage Project and China Scholarship Council. The author also thanks Kunyu Zhao for performing the PFM measurements.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hsu, H.S.; Benjauthrit, V.; Zheng, F.; Chen, R.; Huang, Y.; Zhou, Q.; Shung, K.K. PMN-PT–PZT composite films for high frequency ultrasonic transducer applications. Sensor Actuat. A Phys. 2012, 179, 121–124. [Google Scholar] [CrossRef] [Green Version]
  2. Near, C.; Gentilman, R.; Pazol, B.; Fiore, D.; Serwatka, W.; Guyen, H.P.; Markowski, K.; Mcguire, P.; Bowen, L. Application of piezoelectric composites and net: Hape PZT ceramics as acoustic sensors and actuators. J. Acoust. Soc. Am. 1998, 101, 221–229. [Google Scholar] [CrossRef]
  3. Rathod, V.T. A Review of Electric Impedance Matching Techniques for Piezoelectric Sensors, Actuators and Transducers. Electronics 2019, 8, 169. [Google Scholar]
  4. Roedel, J. ChemInform Abstract: Giant Electric-Field-Induced Strains in Lead-Free Ceramics for Actuator Applications-Status and Perspective. J. Electroceram. 2012, 29, 71–93. [Google Scholar]
  5. Rodel, J.; Jo, W.; Seifert, K.T.P.; Anton, E.M.; Damjanovic, D. Perspective on the Development of Lead-Free Piezoceramics. J. Am. Ceram. Soc. 2009, 92, 1153–1177. [Google Scholar]
  6. Saito, Y.; Takao, H.; Tani, T.; Nonoyama, T.; Takatori, K.; Homma, T.; Nagaya, T.; Nakamura, M. Lead-free piezoceramics. Nature 2004, 432, 84–87. [Google Scholar] [CrossRef]
  7. Shuai, C.; Zhang, B.; Lei, Z.; Wang, K. Enhanced insulating and piezoelectric properties of 0.7BiFeO3–0.3BaTiO3 lead-free ceramics by optimizing calcination temperature: Analysis of Bi3+ volatilization and phase structures. J. Mater. Chem. C. 2018, 6, 3982–3989. [Google Scholar]
  8. Wang, X.; Wu, J.; Xiao, D.; Zhu, J.; Cheng, X.; Zheng, T.; Zhang, B.; Lou, X.; Wang, X. Giant Piezoelectricity in Potassium-Sodium Niobate Lead-Free Ceramics. J. Am. Chem. Soc. 2014, 136, 2905–2910. [Google Scholar]
  9. Kakimoto, K.I.; Akao, K.; Guo, Y.; Ohsato, H. Raman Scattering Study of Piezoelectric (Na0.5K0.5)NbO3-LiNbO3 Ceramics. Jpn. J. Appl. Phys. 2005, 44, 7064–7067. [Google Scholar] [CrossRef]
  10. Takenaka, T.; Maruyama, K.I.; Sakata, K. (Bi1/2Na1/2)TiO3-BaTiO3 system for lead-free piezoelectric ceramics. Jpn. J. Appl. Phys. 1991, 30, 2236–2239. [Google Scholar] [CrossRef]
  11. Cheng, M.; Guo, H.; Beckman, S.P.; Tan, X. Creation and Destruction of Morphotropic Phase Boundaries through Electrical Poling: A Case Study of Lead-Free (Bi1/2Na1/2)TiO3-BaTiO3 Piezoelectrics. Phys. Rev. Lett. 2012, 109, 107601–107602. [Google Scholar]
  12. Ahart, M.; Somayazulu, M.; Cohen, R.E.; Ganesh, P.; Dera, P.; Mao, H.K.; Hemley, R.J.; Ren, Y.; Liermann, P.; Wu, Z. Origin of morphotropic phase boundaries in ferroelectrics. Nature 2008, 451, 545–548. [Google Scholar] [CrossRef] [PubMed]
  13. Daniels, J.E.; Jo, W.; Rödel, J.; Jones, J.L. Electric-field-induced phase transformation at a lead-free morphotropic phase boundary: Case study in a 93%(Bi0.5Na0.5)TiO3-7%BaTiO3 piezoelectric ceramic. Appl. Phys. Lett. 2014. [Google Scholar] [CrossRef] [Green Version]
  14. Ge, W.; Luo, C.; Zhang, Q.; Devreugd, C.P.; Ren, Y.; Li, J.; Luo, H.; Viehland, D. Ultrahigh electromechanical response in (1-x)(Na0.5)Bi0.5)TiO3-xBaTiO3 single-crystals via polarization extension. J.Appl. Phys. 2012, 111, 93501–93508. [Google Scholar] [CrossRef] [Green Version]
  15. Wang, Y.; Luo, C.; Wang, S.; Chen, C.; Yuan, G.; Luo, H.; Viehland, D. Large Piezoelectricity in Ternary Lead-Free Single Crystals. Adv. Electron. Mater. 2020, 6, 1900949. [Google Scholar]
  16. Peng, J.; Luo, H.; He, T.; Xu, H.; Lin, D. Elastic, dielectric, and piezoelectric characterization of 0.70Pb(Mg1/3Nb2/3)O3–0.30PbTiO3 single crystals. Mater. Lett. 2005, 59, 640–643. [Google Scholar] [CrossRef]
  17. Yao, F.Z.; Wang, K.; Li, J.F. Comprehensive investigation of elastic and electrical properties of Li/Ta-modified (K,Na)NbO_3 lead-free piezoceramics. J. Appl. Phys. 2013, 113, 174101–174105. [Google Scholar] [CrossRef]
  18. Huo, X.; Zheng, L.; Zhang, R.; Wang, R.; Wang, J. A high quality lead-free (Li, Ta) modified (K, Na)NbO3 single crystal and its complete set of elastic, dielectric and piezoelectric coefficients with macroscopic 4mm symmetry. Crystengcomm 2014. [Google Scholar] [CrossRef]
  19. Huo, X.; Zhang, R.; Zheng, L.; Zhang, S.; Cao, W. (K, Na, Li)(Nb, Ta)O3:Mn Lead-Free Single Crystal with High Piezoelectric Properties. J. Am. Ceram Soc. 2015, 98, 1829–1835. [Google Scholar] [CrossRef] [Green Version]
  20. Zhou, D.; Lam, K.H.; Chen, Y.; Zhang, Q.; Chiu, Y.C.; Luo, H.; Dai, J.; Chan, H.L.W. Lead-free piezoelectric single crystal based 1-3 composites for ultrasonic transducer applications. Sensor. Actuat. A Phys. 2012, 182, 95–100. [Google Scholar]
  21. Zheng, L.; Yi, X.; Zhang, S.; Jiang, W.; Yang, B.; Zhang, R.; Cao, W. Complete set of material constants of 0.95(Na0.5Bi0.5)TiO3-0.05BaTiO3 lead-free piezoelectric single crystal and the delineation of extrinsic contributions. Appl. Phys. Lett. 2013, 103, 122905. [Google Scholar] [CrossRef] [Green Version]
  22. Ma, C.; Tan, X.; Kleebe, H.J. In situ Transmission Electron Microscopy Study on the Phase Transitionsin Lead-Free (1−x)(Bi1/2Na1/2)TiO3–xBaTiO3 Ceramics. J. Am. Ceram. Soc. 2011, 94, 4040–4044. [Google Scholar] [CrossRef] [Green Version]
  23. Chao, C.; Zhang, H.; Hao, D.; Huang, T.; Li, X.; Zhao, X.; Hu, Z.; Dong, W.; Luo, H. Electric field and temperature-induced phase transition in Mn-doped Na1/2Bi1/2TiO3-5.O at.%BaTiO3 single crystals investigated by micro-Raman scattering. Appl. Phys. Lett. 2014, 104, 142901–142902. [Google Scholar]
  24. Glazer, A.M. The classification of tilted octahedra in perovskites. Acta. Crystallographica 1972, 28, 3384–3392. [Google Scholar] [CrossRef]
  25. Kreisel, J.; Glazer, A.M.; Jones, G.; Thomas, P.A.; Abello, L.; Lucazeau, G. An x-ray diffraction and Raman spectroscopy investigation of A-site substituted perovskite compounds: The (Na1-xKx)0.5Bi0.5TiO3 (0 <x <1) solid solution. J. Phys. Condens. Matter 2000, 12, 3267. [Google Scholar]
  26. Maurya, D.; Pramanick, A.; Feygenson, M.; Neuefeind, J.C.; Priya, S. Effect of poling on nanodomains and nanoscale structure in A-site disordered lead-free piezoelectric Na0.5Bi0.5TiO3-BaTiO3. J. Mater. Chem. C. 2014, 2. [Google Scholar] [CrossRef]
  27. Ge, W.; Hu, C.; Devreugd, C.; Li, J.; Viehl, D. Influence of BaTiO3 Content on the Structure and Properties of Na0.5Bi0.5TiO3 Crystals. J. Am. Ceram. Soc. 2011, 94, 3084–3087. [Google Scholar] [CrossRef]
  28. Yao, J.; Li, Y.; Ge, W.; Liang, L.; Li, J.; Viehland, D.; Zhang, Q.; Luo, H. Evolution of domain structures in Na1/2Bi1/2TiO3 single crystals with BaTiO3. Phys. Rev. B. 2011, 83, 54101–54107. [Google Scholar] [CrossRef] [Green Version]
  29. Rossetti, G.A.; Khachaturyan, A.G.; Akcay, G.; Ni, Y. Ferroelectric solid solutions with morphotropic boundaries: Vanishing polarization anisotropy, adaptive, polar glass, and two-phase states. J. Appl. Phys. 2008, 103, 114113. [Google Scholar] [CrossRef]
  30. Schoenau, K.A.; Schmitt, L.A.; Fuess, H.; Knapp, M.; Hoffmann, M.J. Nanodomain Structure of Pb(Zr1-xTix)O3 at Its Morphotropic Phase Boundary: Investigations From Local to Average Structure. Phys. Rev. B 2007, 75, 184117. [Google Scholar] [CrossRef] [Green Version]
  31. Bai, F.; Li, J.; Viehland, D. Domain hierarchy in annealed (001)-oriented Pb(Mg1/3Nb2/3)O3-x%PbTiO3 single crystals. Appl. Phys. Lett. 2004, 85, 2313–2315. [Google Scholar] [CrossRef] [Green Version]
  32. Wada, S.; Kakemoto, H.; Tsurumi, T. Enhanced Piezoelectric Properties of Piezoelectric Single Crystals by Domain Engineering. Mater. Trans. 2004, 45, 178–187. [Google Scholar] [CrossRef] [Green Version]
  33. Wada, S.; Tsurumi, T. Enhanced piezoelectricity of barium titanate single crystals with engineered domain configuration. Brit. Ceram. Trans. 2004, 103, 93–96. [Google Scholar]
  34. Yako, K.; Kakemoto, H.; Tsurumi, T.; Wada, S. Enhanced Piezoelectric Properties of Barium Titanate Single Crystals by Domain Engineering. Key Eng. Mater. 2006, 301, 23–26. [Google Scholar]
Figure 1. (a) the thickness mode and (b) 31 mode impedance and phase angle of the poled NBT-6BT-KNN crystal along the <001> direction at room temperature; the temperature dependent behavior of (c) kt and (d) k31.
Figure 1. (a) the thickness mode and (b) 31 mode impedance and phase angle of the poled NBT-6BT-KNN crystal along the <001> direction at room temperature; the temperature dependent behavior of (c) kt and (d) k31.
Crystals 10 00435 g001
Figure 2. The temperature dependent dielectric constant and dielectric loss curves of poled <001> oriented NBT-6BT-KNN crystal measured at different frequencies.
Figure 2. The temperature dependent dielectric constant and dielectric loss curves of poled <001> oriented NBT-6BT-KNN crystal measured at different frequencies.
Crystals 10 00435 g002
Figure 3. <001> oriented domain evolution of NBT-6BT-KNN crystal at the following different conditions in sequence (a–d): (a) initial state, electric field E = 0 kV/cm, at room temperature (RT); (b) positive increase of E to 25 kV/cm at RT; (c) remove electric field at RT; (d) heating to 50 °C after removing electric field.
Figure 3. <001> oriented domain evolution of NBT-6BT-KNN crystal at the following different conditions in sequence (a–d): (a) initial state, electric field E = 0 kV/cm, at room temperature (RT); (b) positive increase of E to 25 kV/cm at RT; (c) remove electric field at RT; (d) heating to 50 °C after removing electric field.
Crystals 10 00435 g003
Figure 4. The XRD pattern of unpoled NBT-6BT-KNN crystal powder.
Figure 4. The XRD pattern of unpoled NBT-6BT-KNN crystal powder.
Crystals 10 00435 g004
Figure 5. Raman spectra for <001>-oriented NBT-6BT-KNN crystal at different conditions: (a) room temperature (RT) before poling, (b) RT after poling, (c) 50 °C after poling.
Figure 5. Raman spectra for <001>-oriented NBT-6BT-KNN crystal at different conditions: (a) room temperature (RT) before poling, (b) RT after poling, (c) 50 °C after poling.
Crystals 10 00435 g005
Figure 6. Piezoelectric force microscope (PFM) images of (a,b) NBT-6BT and (c,d) NBT-6BT-KNN crystals.
Figure 6. Piezoelectric force microscope (PFM) images of (a,b) NBT-6BT and (c,d) NBT-6BT-KNN crystals.
Crystals 10 00435 g006
Table 1. Comparison of the properties of previously reported ceramic/single crystals.
Table 1. Comparison of the properties of previously reported ceramic/single crystals.
Matrialsk31ktQuasi-Static d33 (pC/N)Reference
Ceramics
PZT-5H39%60%-[16]
KNN-TL16%34%174[17]
Single crystal
KNN-T46%64.60%162[18]
KNN-TL47%45.10%354[18]
KNN-TL:Mn59%49%630[19]
NBT-0.053BT/epoxy 1-3 composite-73%360[20]
NBT-5BT34.10%54%360[21]
NBT-6BT-KNN45.70%69%840This work

Share and Cite

MDPI and ACS Style

Chen, C.; Yang, L.; Jiang, X.; Huang, X.; Gao, X.; Tu, N.; Shu, K.; Jiang, X.; Zhang, S.; Luo, H. Ultrahigh Electromechanical Coupling and Its Thermal Stability in (Na1/2Bi1/2)TiO3-Based Lead-Free Single Crystals. Crystals 2020, 10, 435. https://doi.org/10.3390/cryst10060435

AMA Style

Chen C, Yang L, Jiang X, Huang X, Gao X, Tu N, Shu K, Jiang X, Zhang S, Luo H. Ultrahigh Electromechanical Coupling and Its Thermal Stability in (Na1/2Bi1/2)TiO3-Based Lead-Free Single Crystals. Crystals. 2020; 10(6):435. https://doi.org/10.3390/cryst10060435

Chicago/Turabian Style

Chen, Chao, Li Yang, Xingan Jiang, Xiaokun Huang, Xiaoyi Gao, Na Tu, Kaizheng Shu, Xiangping Jiang, Shujun Zhang, and Haosu Luo. 2020. "Ultrahigh Electromechanical Coupling and Its Thermal Stability in (Na1/2Bi1/2)TiO3-Based Lead-Free Single Crystals" Crystals 10, no. 6: 435. https://doi.org/10.3390/cryst10060435

APA Style

Chen, C., Yang, L., Jiang, X., Huang, X., Gao, X., Tu, N., Shu, K., Jiang, X., Zhang, S., & Luo, H. (2020). Ultrahigh Electromechanical Coupling and Its Thermal Stability in (Na1/2Bi1/2)TiO3-Based Lead-Free Single Crystals. Crystals, 10(6), 435. https://doi.org/10.3390/cryst10060435

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