In Situ Di-, Piezo-, Ferroelectric Properties and Domain Configurations

Nov 27, 2017 - Key Laboratory of Optoelectronic Materials Chemistry and Physics, Fujian Institute of Research on the Structure of Matter, Chinese Acad...
1 downloads 8 Views 2MB Size
Article Cite This: Cryst. Growth Des. 2018, 18, 145−151

pubs.acs.org/crystal

In Situ Di‑, Piezo‑, Ferroelectric Properties and Domain Configurations of Pb(Sc1/2Nb1/2)O3−Pb(Mg1/3Nb2/3)O3−PbTiO3 Ferroelectric Crystals Zujian Wang, Chao He, Huimin Qiao, Dongfang Pang, Xiaoming Yang, Sangen Zhao, Xiuzhi Li, Ying Liu, and Xifa Long* Key Laboratory of Optoelectronic Materials Chemistry and Physics, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou, Fujian 350002, China S Supporting Information *

ABSTRACT: Relaxor ferroelectric crystals, such as Pb(Mg1/3Nb2/3)O3−PbTiO3 and Pb(Zn1/3Nb2/3)O3−PbTiO3, exhibit high piezoelectric properties. However, they cannot meet the need of high power transducers, due to their low phase transition temperatures and low coercive fields. Thus, it is urgent to develop new-type ferroelectric crystals. Here, a ternary solid solution of Pb(Sc1/2Nb1/2)O3−Pb(Mg1/3Nb2/3)O3−PbTiO3 is a good candidate. By the regulation of growth techniques and the control of raw compositions, Pb(Sc1/2Nb1/2)O3−Pb(Mg1/3Nb2/3)O3−PbTiO3 crystals with different compositions are obtained, which present perovskite phases without any trace of impurity. In situ investigations of di-, piezo-, ferroelectric properties and the observation of domain configurations are realized in Pb(Sc1/2Nb1/2)O3− Pb(Mg1/3Nb2/3)O3−PbTiO3 crystals, which exhibit large di-, ferro-, piezoelectric responses, typical domain switching, high poled and thermal stabilities. At room temperature, the maximum piezoelectric constant is found to be 1550 pC/N, with the maximum peak to peak strain value of 0.57% and coercive field of 4−6 kV/cm. The speckle-shaped nanodomains and butterfly hysteresis loops represented by the variation of amplitudes further indicate large piezoelectric responses of PSN−PMN−PT crystals. Furthermore, the rhombohedral−tetragonal phase transition temperature is more than 120 °C, indicating good thermal stabilities with stable ferroelectricity and large piezoelectric response in a wide temperature range.



INTRODUCTION Relaxor-based ferroelectric solid solutions, such as Pb(Mg1/3Nb2/3)O3−PbTiO3 (PMN−PT) and Pb(Zn1/3Nb2/3)O3−PbTiO3 (PZN−PT) crystals, have attracted much attention, due to their excellent piezoelectric and electromechanical properties. Compared with other ferroelectrics, the critical signature of relaxor-based systems is the existence of polar nanoregions (PNRs) and nanoscale inhomogeneities, coexisting with normal ferroelectric domains, which make a great contribution to large piezoelectric responses of relaxor ferroelectrics.1−6 Furthermore, for high piezoelectric properties, it is important to find the morphotropic phase boundary (MPB) region of relaxor-based ferroelectric solid solution, near which large piezoelectric and dielectric properties are usually found. With compositions (PT mol % ≈ 0.30−0.35) in the vicinity of the MPB region, PMN−PT crystals exhibit large piezoelectric constants (d33 > 2000 pC/N) and high electromechanical coupling factors (k33 > 90%), which are promising candidates for applications in transducers, sensors, and actuators.7−9 However, their relatively low Curie temperature (TC ≈ 130−160 °C), rhombohedral−tetragonal phase transition temperature (Tr‑t ≈ 80−95 °C), and coercive field (Ec ≈ 2−3 kV/cm) greatly restrict © 2017 American Chemical Society

their potential in high temperature and high power applications.9−12 Thus, it is urgent to develop new ferroelectrics with both high stable piezoelectric properties and high thermal stabilities. Ternary solid solutions by doping of a third compound into the PMN−PT system could not only effectively enhance their Curie temperature and coercive field of PMN−PT system, but also maintain good piezoelectric properties. For example, the Pb(In1/2Nb1/2)O3−Pb(Mg1/3Nb2/3)O3−PbTiO3 (PIN−PMN− PT) ternary system by doping of Pb(In1/2Nb1/2)O3 into PMN− PT has been reported with TC > 190 °C, Ec ≈ 5.5 kV/cm, while the Pb(Lu1/2Nb1/2)O3−Pb(Mg1/3Nb2/3O3)−PbTiO3 (PLN− PMN-PT) system exhibits a higher Curie temperature (∼360 °C) and larger coercive field (∼13.8 kV/cm).13−17 Just like PIN and PLN, Pb(Sc1/2Nb1/2)O3 (PSN) is one of the 1:1-type perovskite compounds (Pb(B’B’’)O3), which is also one good doping candidate. The Curie temperature of PSN is about 70 °C,18 higher than that of PMN (about −10 °C). Meanwhile, PSN is ferroelectric at room temperature with a coercive field of 2.2 Received: July 22, 2017 Revised: November 13, 2017 Published: November 27, 2017 145

DOI: 10.1021/acs.cgd.7b01023 Cryst. Growth Des. 2018, 18, 145−151

Crystal Growth & Design

Article

axis diffractometer equipped with Cu Kα radiation and a graphite monochromator was employed to collect powder X-ray diffraction (XRD) patterns at room temperature. The dielectric responses were measured by a computer-controlled Alpha-A broadband dielectric/ impedance spectrometer (NovoControl, GmbH), which were carried out at an ac signal of 1 V (peak to peak) as a function of temperature upon heating from −50 to 400 °C. An aix-ACCT TF analyzer 2000 ferroelectric test system was used to display the P−E and strain-electric field (S−E) loops. The piezoelectric constants were measured by a quasi-static ZJ-4AN d33 meter (Institute of Acoustics, Chinese Academy of Sciences). And the domain configurations at room temperature were observed by Asylum Research ES piezo-response force microscopy (PFM) with a working principle of the inverse piezoelectric effect. Among all methods of domain investigation, PFM can provide nondestructive configurations of ferroelectric domains in a nanometer scale, which are embodied in the form of amplitudes and phases, corresponding to intensities and orientations of domains, respectively. The domain structures of PSN−PMN−PT crystals displayed by amplitude and phase images are accordant, which are not disturbed by the surface morphology (as shown in Figure S1). For PFM observation, the samples were cut into slices with (001)c faces polished to optical standard, which were first annealed at 300 °C for 1 h to release the surface stresses resulting from the cutting and polishing processes. After PFM measurement, the samples were smeared with silver paste as electrodes for the investigation of electrical properties.

kV/cm. Therefore, the doping of PSN into PMN−PT could improve the thermal stability of the PMN−PT system, while keeping highly stable piezoelectric properties. Regarding the Pb(Sc 1/2 Nb 1/2 )O 3 −Pb(Mg1/3 Nb 2/3 )O 3 − PbTiO3 (PSN−PMN-PT) ternary solid solution, related work is very limited. The Yamashita laboratory reported PSN−PMN− PT crystals twice without reference to electrical properties, while the Luo research group obtained 0.05PSN−0.63PMN−0.32PT crystal by the Bridgman technique, with d33 ≈ 1200 pC/N, TC ≈ 160 °C, Tr‑t ≈ 60 °C, which was not improved compared with PMN−PT crystals.19−21 As one new material, PSN−PMN−PT crystals were initially investigated in our previous work, which presented high phase transition temperatures and stable piezoelectric properties.22 For example, 0.39PSN−0.13PMN− 0.48PT crystals with one MPB composition were found to exhibit high TC and Tr‑t of about 230 and 199 °C, respectively, indicating high thermal stability. At the same time, they manifest an enhanced coercive field (about 5.6 kV/cm) with welldeveloped polarization−electric field (P−E) loops, demonstrating stable piezoelectric properties. Therefore, it is necessary to study the PSN−PMN−PT system in the round. In this work, by regulation of growth techniques and control of raw compositions, we designed and grew PSN−PMN−PT crystals with MPB compositions, which exhibit good electrical properties, high poled and thermal stabilities.





RESULTS AND DISCUSSION The XRD patterns of PSN−PMN−PT crystals at room temperature are shown in Figure 2 with the mole ratio 1:3, 1:1, and 3:1 of PSN and PMN, all of which present typical perovskite structures. As the increasing PT content, the structure of PSN− PMN−PT crystals takes a continuous change from the rhombohedral phase to MPB then to the tetragonal phase. For example, as shown in Figure 2a, the single (200) diffraction peak broadens and then splits into two peaks with the increasing PT content, while other peaks of high angles broaden and split into multiple peaks.26−28 Thus, crystals with MPB compositions can be obtained from the nominal composition x (PT mol %) = 0.48−0.50, 0.45−0.47, and 0.47−0.49 for the mole ratio of PSN and PMN of 1:3, 1:1, and 3:1, respectively. The obtained MPB regions are above the line between MPB compositions of PMN− PT and PSN−PT systems, due to the segregation of ions, as shown in Figure 1. With MPB compositions, the (200) diffraction peaks of PSN−PMN−PT crystals can be fitted into three peaks by Gaussian convolution, just as the insets of Figure 2, which is typical MPB characterization with the coexistence of tetragonal and rhombohedral phases (peaks 1 and 3 correspond to the tetragonal phase, while peak 2 corresponds to the rhombohedral phase).

EXPERIMENTAL SECTION

PSN−PMN−PT crystals with different compositions were obtained by the top seeded solution growth (TSSG) technique. The compositions were selected following the specific mole ratio (1:3, 1:1, 3:1) of PSN and PMN in the ternary phase diagram, as shown in Figure 1. A Rigaku R-

Figure 1. Schematic ternary phase diagram of PSN−PMN−PT solid solution with “○” the nominal compositions along different mole ratios of PSN and PMN, and the two dot lines connect the MPB compositions of the PMN−PT and PSN−PT systems.9−12,23−25

Figure 2. XRD patterns of PSN−PMN−PT crystals with different mole ratios of PSN and PMN (a: 1:3; b: 1:1; c: 3:1), the insets of a, b, and c corresponding to the (200) diffraction peaks and Gaussian fitting peaks of 0.1275PSN−0.3825PMN−0.49PT, 0.27PSN−0.27PMN−0.46PT, and 0.39PSN−0.13PMN−0.48PT crystals, respectively 146

DOI: 10.1021/acs.cgd.7b01023 Cryst. Growth Des. 2018, 18, 145−151

Crystal Growth & Design

Article

Figure 3. (a−i) Permittivities of PSN−PMN−PT crystals with compositions near MPB regions as a function of temperature at different frequencies.

As we all know one signature of the ferroelectric phase transition is a prominent dielectric anomaly, which is indeed observed in the PSN−PMN−PT system. The permittivity (ε′) and dielectric loss (tan δ) of [001]-oriented PSN−PMN−PT crystal plates with MPB compositions were measured as a function of temperature at frequencies of 1 Hz, 10 Hz, 100 Hz, and 1 kHz, as shown in Figures 3 and 4. With the increasing frequency, the permittivity decreases, resulting from the lag between frequencies of oscillating dipoles and applied electric field. All of ε′−T curves exhibit two pronounced anomalous permittivity peaks. Upon heating, the first anomaly corresponds to the rhombohedral−tetragonal phase transition, typical MPB behavior of ferroelectric solid solutions,29−31 while the second one associates with ferroelectric−paraelectric phase transition. The rhombohedral−tetragonal phase transition temperature is more than 120 °C, as shown in Figure 5a, which is enhanced compared to that of PMN−PT crystals.32 Hence, PSN−PMN− PT crystals could satisfy applications with a wider temperature range than PMN−PT.33,34 The permittivities at room temperature (ε′RT) are 1800, 1700, and 2000 of 0.1275PSN− 0.3825PMN−0.49PT, 0.27PSN−0.27PMN−0.46PT, and 0.39PSN−0.13PMN−0.48PT crystals, respectively. And the dielectric losses of PSN−PMN−PT crystals are almost less than 5%. Moreover, the electromechanical coupling factors (k33) were calculated to be 77%, 76%, and 74% for [001]-oriented 0.1275PSN−0.3825PMN−0.49PT, 0.27PSN−0.27PMN− 0.46PT, and 0.39PSN−0.13PMN−0.48PT crystals, respectively,

by resonance and antiresonance frequencies (as shown in Figure S2), which is slightly lower than that of PMN−PT crystals.32 Typical perovskite structures and striking dielectric anomalies indicate the ferroelectric characterization of PSN−PMN-PT crystals, which was further embodied by well-developed P−E hysteresis loops. As shown in Figure 6a,c,e, the ferroelectric hysteresis loops of [001]-oriented PSN−PMN−PT crystal plates are displayed by excellent P−E rectangle at the frequency of 2 Hz at different temperature, exhibiting typical domain switching, which is also identified in microscopic domain configurations under an external electric field (as shown in Figure S3). At room temperature, by a bipolar electric field of about ±20 kV/cm, the polarization reaches its saturation, with the remnant polarization (Pr) and coercive field more than 20 μC/cm2 and 4 kV/cm, respectively, as shown in Figure 5b−d. The coercive field is twice of that of PMN−PT crystals (as listed in Table 1), indicating stable piezoelectric properties.11 The loops of other temperatures are also investigated. Upon heating, the loops become slimmer and slimmer with a smaller and smaller coercive field, which indicates that the polarization reversal becomes more and more easy. The remnant polarization changes a little as the increasing temperature for 0.27PSN−0.27PMN−0.46PT and 0.39PSN−0.13PMN−0.48PT crystals, while it varies relatively much for 0.1275PSN−0.3875PMN−0.49PT crystal, which increases from 28.5 μC/cm2 to the maximum value of 32.5 μC/cm2 at the temperature of 125 °C, due to multiple polarization directions resulting from temperature induced rhombohedral−tetragonal phase transition. In any event, 147

DOI: 10.1021/acs.cgd.7b01023 Cryst. Growth Des. 2018, 18, 145−151

Crystal Growth & Design

Article

Figure 4. (a−i) Dielectric losses of PSN−PMN−PT crystals with different compositions as a function of temperature at different frequencies.

Figure 5. Rhombohedral−tetragonal phase transition temperature of different compositions (a); the coercive field and remnant polarization as a function of temperature (b: 0.1275PSN−0.3825PMN−0.49PT crystal; c: 0.27PSN−0.27PMN−0.46PT crystal; d: 0.39PSN−0.13PMN−0.48PT crystal).

PSN−PMN−PT crystals exhibit good thermal stability at the aspect of ferroelectric properties.

The good piezoelectric properties of PSN−PMN−PT crystals reflect in the excellent strain−electric field (S−E) loops, as 148

DOI: 10.1021/acs.cgd.7b01023 Cryst. Growth Des. 2018, 18, 145−151

Crystal Growth & Design

Article

Figure 6. Ferroelectric hysteresis loops and butterfly strain curves as a function of electric field at different temperatures (a, b: 0.1275PSN− 0.3825PMN−0.49PT crystal; c, d: 0.27PSN−0.27PMN−0.46PT crystal; e, f: 0.39PSN−0.13PMN−0.48PT crystal).

Table 1. Electrical Properties of Relaxor-Based Ferroelectric Crystals crystal 7−9,19−21

PMN−PT PIN−PMN−PT37 PIN−PMN−PT38 PIN−PMN−PT39 PIN−PMN−PT40 PIN−PMN−PT15 PSN−PMN−PTb a

Tr‑t (°C)

TC (°C)

Eca (kV/cm)

d33a (pC/N)

ε′RT

k33 (%)

80−95 90−110 ∼115 110−132 ∼120 ∼150 120−180

130−160 125−175 ∼135 164−218 ∼185 190−200 200−240

2−3

>2000 1100−1800 1000−2100 550−2800 ∼2200 2200−2600 1260−1550

3000 3500−6000 ∼2500 ∼4000 ∼5000 1500−2500 1700−2000

>90

∼3.3 6−8 ∼5.7 ∼5.5 4−6

>80 ∼92 ∼90 ∼75

At room temperature. bThis work.

which are almost consistent with the measured results. This large piezoelectric constant may result from the composition induced rhombohedral−tetragonal phase transition, which leads to multiple polarization directions. In this instable polarization state, polarization switch is relatively easy by external stress or electric field. Furthermore, PNRs play an important role in large piezoelectric responses, which have been observed in the speckleshaped domain configurations with size of 8−700 nm (as shown in Figure S3), exhibiting relaxor characteristics of PSN−PMN− PT crystals.35,36 Before the rhombohedral−teragonal phase transition temperature, the piezoelectric constants become larger and larger with the increasing temperature, due to more active domains and lower coercive fields as the increasing temperature. The maximum piezoelectric constants are calculated to be 2600

shown in Figure 6b,d,f and Figure 7. Butterfly-like S−E loops manifest piezoelectric responses under a bipolar drive, which are also identified in microscopic domain configurations under a bias voltage of about 20 V (as shown in Figure S3). The minimum peak to peak strain value at room temperature is 0.28%, more than that (0.12%) of PMN−0.30PT crystal.10 At the same time, the elliptic S−E loops with large angle of inclination indicate large piezoelectric responses, from which piezoelectric constants can be obtained. The calculated piezoelectric constants are displayed as a function of temperature, as shown in Figure 6d. At room temperature, the values of the piezoelectric constant are obtained to be 1550 pC/N, 1280 pC/N, and 1260 pC/N for 0.1275PSN−0.3875PMN−0.49PT, 0.27PSN−0.27PMN− 0.46PT, and 0.39PSN−0.13PMN−0.48PT crystals, respectively, 149

DOI: 10.1021/acs.cgd.7b01023 Cryst. Growth Des. 2018, 18, 145−151

Crystal Growth & Design

Article

Figure 7. (a−d) Unipolar strain curves and the calculated piezoelectric constant at different temperatures.



pC/N (at 125 °C), 1900 pC/N (at 150 °C), and 2760 pC/N (at 150 °C) for 0.1275PSN−0.3825PMN−0.49PT, 0.27PSN− 0.27PMN−0.46PT, and 0.39PSN−0.13PMN-0.48PT crystals, respectively, so we can see that PSN−PMN−PT crystals always exhibit large piezoelectric response in a wide temperature range.



*E-mail: [email protected]. ORCID

Xifa Long: 0000-0003-2960-9461 Notes

CONCLUSION

The authors declare no competing financial interest.



By the regulation and control of the composition, PSN−PMN− PT crystals with MPB compositions are designed and obtained. The perovskite structures and MPB regions of PSN−PMN−PT crystals are confirmed by structure analyses. Especially, an in situ study of di-, ferro-, piezoelectric properties as a function of temperature and the observation of domain switching exhibit their good electrical properties, high thermal and poled stabilities. Compared with PMN−PT crystals, PSN−PMN−PT crystals manifest higher phase transition temperatures and larger coercive fields, which are almost equivalent with those of the state-of-the-art PIN−PMN−PT crystals (as listed in Table 1). However, their piezoelectric constants and electromechanical coupling factors are somewhat less than those of PMN−PT and PIN−PMN−PT crystals, which can be improved by the optimization of crystal composition and quality. In general, PSN−PMN−PT crystals are promising ferroelectrics, which are good candidates for high temperature and high power applications.



AUTHOR INFORMATION

Corresponding Author

ACKNOWLEDGMENTS This work was supported by the National Natural Science Foundation of China under Grant Nos. 91422303, 11504373, 11404331, and 5160230, the Science and Technology Project of Fujian Province (2015H0049), the Strategic Priority Research Program of the Chinese Academy of Sciences (XDB20000000).



REFERENCES

(1) Manley, M. E.; Abernathy, D. L.; Sahul, R.; Parshall, D. E.; Lynn, J. W.; Christianson, A. D.; Stonaha, P. J.; Specht, E. D.; Budai, J. D. Sci. Adv. 2016, 2, e1501814. (2) Hu, B. B.; Zhu, M. K.; Guo, J. J.; Wang, Y.; Zheng, M. P.; Hou, Y. D. J. Am. Ceram. Soc. 2016, 99, 1637−1644. (3) Li, F.; Zhang, S. J.; Xu, Z.; Chen, L.-Q. Adv. Funct. Mater. 2017, 27, 1700310. (4) Li, F.; Zhang, S. J.; Yang, T. T.; Xu, Z.; Zhang, N.; Liu, G.; Wang, J. J.; Wang, J. L.; Cheng, Z. X.; Ye, Z.-G.; Luo, J.; Shrout, T. R.; Chen, L.-Q. Nat. Commun. 2016, 7, 13807. (5) Hagiwara, M.; Ehara, T.; Novak, N.; Khansur, N. H.; Ayrikyan, A.; webber, K. G.; Fujihara, S. Phys. Rev. B: Condens. Matter Mater. Phys. 2017, 96, 014103. (6) Helal, M. A.; Aftabuzzaman, M.; Tsukada, S.; Kojima, S. Sci. Rep. 2017, 7, 44448. (7) Park, S.-E.; Shrout, T. R. J. Appl. Phys. 1997, 82, 1804−1811. (8) Service, R. F. Science 1997, 275, 1878. (9) Viehland, D.; Powers, J. Appl. Phys. Lett. 2001, 78, 3112−3114. (10) Long, X. F.; Ye, Z.-G. Acta Mater. 2007, 55, 6507−6512. (11) Ye, Z.-G. MRS Bull. 2009, 34, 277−283. (12) Feng, Z. Y.; Zhao, X. Y.; Luo, H. S. J. Appl. Phys. 2006, 100, 024104.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.cgd.7b01023. Morphology and domain configurations, Resonance and antiresonance frequencies, Piezoelectric, ferroelectric responses and domain sizes (PDF) 150

DOI: 10.1021/acs.cgd.7b01023 Cryst. Growth Des. 2018, 18, 145−151

Crystal Growth & Design

Article

(13) Chang, Y. F.; Wu, J.; Sun, Y.; Zhang, S. T.; Wang, X. H.; Yang, B.; Messing, G. L.; Cao, W. W. Appl. Phys. Lett. 2015, 107, 082902. (14) Tian, J.; Han, P. d.; Huang, X. L.; Pan, H. X.; Carroll, J. F.; Payne, D. A. Appl. Phys. Lett. 2007, 91, 222903. (15) Li, X. Z.; Wang, Z. J.; He, C.; Long, X. F.; Ye, Z.-G. J. Appl. Phys. 2012, 111, 034105. (16) Chen, J. W.; Li, X. B.; Zhao, X. Y.; Zhang, H. W.; Chen, C.; Wang, X. A.; Ren, B.; Di, W. N.; Deng, H.; Luo, H. J. Cryst. Growth 2015, 423, 50−54. (17) Liu, Y.; Lai, F. C.; Huang, Z. G.; Shen, D. Q.; Long, X. F. J. Inorg. Mater. 2014, 29, 912−916. (18) Wang, Z. J.; Long, X. F. CrystEngComm 2014, 16, 6588−6592. (19) Hosono, Y.; Harada, K.; Yamashita, Y.; Dong, M.; Ye, Z.-G. Jpn. J. Appl. Phys. 2000, 39, 5589−5592. (20) Hosono, Y.; Yamashita, Y.; Sakamoto, H.; Ichinose, N. Jpn. J. Appl. Phys. 2003, 42, 6062−6067. (21) Guo, Y. P.; Xu, H. Q.; Luo, H. S.; Xu, G. S.; Yin, Z. W. J. Cryst. Growth 2001, 226, 111−116. (22) Wang, Z. J.; He, C.; Li, X. Z.; Liu, Y.; Long, X. F.; Han, S. J.; Pan, S. L. Mater. Lett. 2016, 184, 162−165. (23) Xia, Z. G.; Li, Q. Solid State Commun. 2007, 142, 323−328. (24) Fan, H. Q.; Kim, H.-E. J. Am. Ceram. Soc. 2001, 84, 636−638. (25) Song, F. B.; Li, Q.; Zhong, H. S.; Li, C. H.; Zhao, S. X.; Shen, D. Z. Mater. Chem. Phys. 2004, 83, 135−139. (26) Bing, Y.-H.; Ye, Z.-G. J. Cryst. Growth 2006, 287, 326−329. (27) Long, X. F.; Ling, J. B.; Li, X. Z.; Wang, Z. J.; Ye, Z.-G. Cryst. Growth Des. 2009, 9, 657−659. (28) He, C.; Wang, Z. J.; Li, X. Z.; Liu, Y.; Shen, D. Q.; Li, T.; Long, X. F. J. Phys. D: Appl. Phys. 2012, 45, 105305. (29) Kuwata, J.; Uchino, K.; Nomura, S. Jpn. J. Appl. Phys. 1982, 21, 1298−1302. (30) Zhang, S. J.; Randall, C. A.; Shrout, T. R. Appl. Phys. Lett. 2003, 83, 3150−3152. (31) Zhang, S. J.; Xia, R.; Lebrun, L.; Anderson, D.; Shrout, T. R. Mater. Lett. 2005, 59, 3471−3475. (32) Park, S.-E.; Shrout, T. R. IEEE Ultrason. Sympo. 1996, 935−942. (33) Fu, H. X.; Cohen, R. E. Nature 2000, 403, 281−283. (34) Noheda, B.; Cox, D. E.; Shirane, G.; Park, S.-E.; Cross, L. E.; Zhong, Z. Phys. Rev. Lett. 2001, 86, 3891−3894. (35) He, W. H.; Jiang, T.; Li, Q.; Xi, X. Q.; Yan, Q. F. J. Am. Ceram. Soc. 2017, 100, 1724−1732. (36) Liu, D.; Ma, C. G.; Luo, H. S.; Tang, Y. X.; Shi, W. Z.; Sun, D. Z.; Wang, F. F.; Wang, T. J. Alloys Compd. 2017, 696, 166−170. (37) Li, Z. R.; Song, K. X.; Guo, H. S.; Liu, Y. B.; Ma, M.; Fan, S. J.; Xu, Z. J. Cryst. Growth 2017, 468, 331−334. (38) Wan, Y. H.; Li, Z. R.; Xu, Z.; Fan, S. J.; Yao, X. J. Alloys Compd. 2013, 558, 244−247. (39) Chen, J. W.; Li, X. B.; Zhao, X. Y.; Wang, X. A.; Chen, C.; Deng, H.; Ren, B.; Jiao, J.; Luo, H. S. J. Mater. Sci.: Mater. Electron. 2015, 26, 9316−9328. (40) Wang, X.; Lin, D.; Wang, S.; Chen, J. W.; Xu, H. Q.; Li, X. B.; Zhao, X. Y.; Luo, H. S. J. Cryst. Growth 2016, 452, 105−110.

151

DOI: 10.1021/acs.cgd.7b01023 Cryst. Growth Des. 2018, 18, 145−151