Favorable dynamics of switching and non-switching polarizations in ⟨001⟩pc oriented Sm:PMN-PT crystal

Perovskite (1−x)Pb(Mg1/3Nb2/3)O3−xPbTiO3 (PMN−xPT) is a well-known relaxor with excellent piezoresponse in the composition range of 0.29 11. S.-E. Park and T. R. Shrout, J. Appl. Phys. 82(4), 1804 (1997). https://doi.org/10.1063/1.365983 The high piezoelectric activity is often attributed to the presence of a broad morphotropic phase boundary (MPB) between the rhombohedral (R) and tetragonal (T) crystallographic phases in this composition range.11. S.-E. Park and T. R. Shrout, J. Appl. Phys. 82(4), 1804 (1997). https://doi.org/10.1063/1.365983 A few crystallographic investigations have also suggested a role of the low symmetry monoclinic (Mc) phase in this high functional activity.22. B. Noheda, D. E. Cox, G. Shirane, J. Gao, and Z. G. Ye, Phys. Rev. B 66(5), 054104 (2002); https://doi.org/10.1103/PhysRevB.66.054104 B. Noheda, Curr. Opin. Solid State Mater. Sci. 6(1), 27 (2002). https://doi.org/10.1016/S1359-0286(02)00015-3 At x ≤ 0.29, PMN-xPT exhibits R structure with comparatively weak piezoresponse. However, recently, Li et al.3,43. F. Li, D. Lin, Z. Chen, Z. Cheng, J. Wang, C. Li, Z. Xu, Q. Huang, X. Liao, L.-Q. Chen, T. R. Shrout, and S. Zhang, Nat. Mater. 17, 349 (2018). https://doi.org/10.1038/s41563-018-0034-44. F. Li, M. J. Cabral, B. Xu, Z. Cheng, E. C. Dickey, J. M. LeBeau, J. Wang, J. Luo, S. Taylor, W. Hackenberger, L. Bellaiche, Z. Xu, L.-Q. Chen, T. R. Shrout, and S. Zhang, Science 364(6437), 264 (2019). https://doi.org/10.1126/science.aaw2781 have shown that a small fraction of samarium doping to x = 0.29 ceramic and 001pc-oriented crystal enhances their respective piezoelectric activity significantly. It was suggested that the interfacial energy, arising due to the presence of orthorhombic (O) nanodomains in the predominant T matrix, helps in achieving the large longitudinal piezoresponse through the ease of polarization realignment under the applied electric field.A follow-up crystallographic investigation by Zao et al.55. C. Zhao, F. Li, S. Zhang, S. Li, and J. L. Jones, J. Appl. Phys. 126(7), 075101 (2019). https://doi.org/10.1063/1.5089477 revealed that the maximum piezoelectric strain in tetragonal 0.71(Pb0.9625Sm0.025)(Mg1/3Nb2/3)O3-0.29PbTiO3 originates from the crystallites having ⟨001⟩pc parallel to the electric field, suggesting the polarization extension to be the mechanism responsible for the high piezoresponse rather than the ease of its realignment.66. M. Davis, M. Budimir, D. Damjanovic, and N. Setter, J. Appl. Phys. 101(5), 054112 (2007). https://doi.org/10.1063/1.2653925 However, it was also suggested that the local R disorder in the samarium doped PMN-0.29PT ceramic may result in the realignment of polarization along ⟨001⟩pc (R → T) under the applied electric field to facilitate the high piezoelectricity. Undoped PMN-0.28PT ceramic and crystal have also been reported previously to exhibit a similar R → T phase transition under the applied field; however, this polarization rotation does not accompany any noticeable piezoelectric strain.77. A. A. Levin, A. I. Pommrich, T. Weißbach, P. Paufler, and D. C. Meyer, Int. J. Mater. Res. 98(8), 700 (2007); https://doi.org/10.3139/146.101521 K. Chen, X.-W. Zhang, and H. Luo, J. Phys.: Condens. Matter 14, L571 (2002). https://doi.org/10.1088/0953-8984/14/29/103 In our recent investigation on PMN-0.28PT, we have shown that only a marginal enhancement in its piezoelectric activity can be achieved by means of improving the domain switching.88. S. Gupta and R. L. Mahajan, Comput. Mater. Sci. 200, 110798 (2021). https://doi.org/10.1016/j.commatsci.2021.110798 Given the diversity of mechanisms proposed, there is a need to further investigate the origin of the giant piezoelectric activity in the samarium doped PMN-0.28PT crystal and polycrystalline ceramic.In this investigation, we perform a comparative dynamic scaling analysis on ⟨001⟩pc oriented PMN-PT crystal and its samarium doped counterpart to understand their distinct responses under the electric stimulus. In the conventional dynamic scaling approach, the area under the ferroelectric (FE) hysteresis loops (⟨A⟩) is studied as a function of amplitude (E) and frequency (f) of the poling field.9–119. S. Gupta and S. Priya, Appl. Phys. Lett. 98(24), 242906 (2011). https://doi.org/10.1063/1.360005810. R. Yimnirun, Y. Laosiritaworn, S. Wongsaenmai, and S. Ananta, Appl. Phys. Lett. 89(16), 162901 (2006). https://doi.org/10.1063/1.236314311. Y. Yang, E. Sun, H. Zheng, B. Yang, R. Zhang, and W. Cao, Appl. Phys. Lett. 119(18), 182902 (2021). https://doi.org/10.1063/5.0067955 In general, ⟨A⟩ represents the energy dissipated in the associated domain reversal process and is related to the two parameters according to the power law relation ⟨A⟩∝ Eαfβ. Here, the parameters α and β reveal the ability of crystallographic domains to respond to the changing amplitude and frequency of the electric stimulus, respectively, and have distinct values for different materials. A number of investigations have been undertaken in the past to utilize this approach for single crystals,9,11,129. S. Gupta and S. Priya, Appl. Phys. Lett. 98(24), 242906 (2011). https://doi.org/10.1063/1.360005811. Y. Yang, E. Sun, H. Zheng, B. Yang, R. Zhang, and W. Cao, Appl. Phys. Lett. 119(18), 182902 (2021). https://doi.org/10.1063/5.006795512. N. Wongdamnern, A. Ngamjarurojana, Y. Laosiritaworn, S. Ananta, and R. Yimnirun, J. Appl. Phys. 105(4), 044109 (2009). https://doi.org/10.1063/1.3086317 polycrystalline ceramics,10,1310. R. Yimnirun, Y. Laosiritaworn, S. Wongsaenmai, and S. Ananta, Appl. Phys. Lett. 89(16), 162901 (2006). https://doi.org/10.1063/1.236314313. S. Gupta and S. Priya, Appl. Phys. Lett. 102(1), 012906 (2013); https://doi.org/10.1063/1.4773983 N. Wongdamnern, K. Kanchiang, A. Ngamjarurojana, S. Ananta, Y. Laosiritaworn, A. Charoenphakdee, S. Gupta, S. Priya, and R. Yimnirun, Smart Mater. Struct. 23(8), 085022 (2014). https://doi.org/10.1088/0964-1726/23/8/085022 and thin films.1414. B. Pan, H. Yu, D. Wu, X. H. Zhou, and J.-M. Liu, Appl. Phys. Lett. 83(7), 1406 (2003). https://doi.org/10.1063/1.1602580 However, as the area enclosed by a ferroelectric loop also has a non-switching contribution to it, the response cannot be attributed to the domain reorientation process only. With this in mind, we undertook to perform the dynamic scaling analysis on switching and non-switching polarization loops separately. This approach not only provides a more accurate picture of the domain reversal process but also facilitates an opportunity to analyze the dynamics of the non-switching intrinsic response separately. Revealing the dynamics of non-switching polarization is equally important in order to understand the mechanism of high piezoelectricity in the samarium doped PMN-xPT crystal.Two crystals used in this investigation were grown by the molten salt method using PbO (having 10% wt./wt. B2O3) as flux.1515. T. R. Shrout, Z. P. Chang, N. Kim, and S. Markgraf, Ferroelectr. Lett. Sect. 12(3), 63 (1990). https://doi.org/10.1080/07315179008201118 The ceramic powders of compositions 0.72Pb(Mg1/3Nb2/3)O3-0.28PbTiO3 and 0.72(Pb0.985Sm0.01)(Mg1/3Nb2/3)O3-0.28PbTiO3 were calcined by the solid-state synthesis route using high purity (>99.5%) oxide precursors.33. F. Li, D. Lin, Z. Chen, Z. Cheng, J. Wang, C. Li, Z. Xu, Q. Huang, X. Liao, L.-Q. Chen, T. R. Shrout, and S. Zhang, Nat. Mater. 17, 349 (2018). https://doi.org/10.1038/s41563-018-0034-4 A controlled cooling of the molten solution of flux and ceramic from 1300 to 1000 °C resulted in the growth of cuboid-shaped crystals of lateral dimensions of few millimeters. Less than unity segregation coefficient of PbTiO3 and poor incorporation of samarium in PMN-PT crystals4,164. F. Li, M. J. Cabral, B. Xu, Z. Cheng, E. C. Dickey, J. M. LeBeau, J. Wang, J. Luo, S. Taylor, W. Hackenberger, L. Bellaiche, Z. Xu, L.-Q. Chen, T. R. Shrout, and S. Zhang, Science 364(6437), 264 (2019). https://doi.org/10.1126/science.aaw278116. I. Bhaumik, G. Singh, S. Ganesamoorthy, A. K. Karnal, M. K. Tiwari, and V. S. Tiwari, Cryst. Res. Technol. 42(4), 356 (2007). https://doi.org/10.1002/crat.200610828 confirm the R nature of both the crystals (from here onwards abbreviated as PMN-PT and Sm:PMN-PT, respectively), which is supported by the dielectric constant vs temperature data (Fig. S1 of the supplementary material) as well. The compositions of the two crystals have been chosen in such a way that any distinction in their dynamic scaling behavior can be squarely attributed to the effect of samarium doping, rather than any crystal structure dependent dissimilarity in domain configuration and their distribution with respect to the poling field along ⟨001⟩pc. For electric measurements, crystals were coated with silver paint on the major facets, followed by firing at 650 °C. An electric poling at a field of 15 kV/cm resulted in low field longitudinal piezoelectric coefficient (d33) values of 1000–1100 and 1800–1900 pC/N for PMN-PT and Sm:PMN-PT, respectively, confirming their distinct piezoelectric behavior.To perform the dynamic scaling analysis on the two crystals, they were subjected to the switching/non-switching polarization measurements as a function of frequency and amplitude of the applied poling field. In the process, the amplitude of the field was varied from 1 to 15 kV/cm, whereas a frequency range of 1–200 Hz was covered. Figure 1 illustrates the representative sequence of the electric field pulses employed to measure the switching/non-switching polarization loops in comparison with the one employed for conventional ferroelectric measurements. For the measurement of conventional ferroelectric loop, a continuous full triangular wave following a preset pulse is employed. On the other hand, to measure the switching and non-switching polarization loops separately, a preset pulse followed by two successive half waves of two polarities [Fig. 1(b)] are applied to the sample.1717. M. Fukunaga and Y. Noda, J. Phys. Soc. of Japan 77(6), 064706 (2008). https://doi.org/10.1143/JPSJ.77.064706 Once the preset pulse sets the domain configuration to a given state, the first triangular half wave of opposite polarity causes a domain reversal, and the response of the material has both intrinsic (non-switching) and extrinsic (switching) contributions to it. For the subsequent half wave of the same polarity, no domain reversal takes place, and the response of the crystal has only non-switching contribution. By subtracting the response obtained for the second half wave from that for the first one, switching response of the crystal can be determined. Similarly, the next two half waves give the switching/non-switching responses of the crystal for the opposite polarity of the electric field.Figures 2(a) and 2(b) exhibit the switching (SW), non-switching (NSW), and ferroelectric (FE = SW + NSW) polarization loops for PMN-PT and Sm:PMN-PT crystals obtained at 15 kV/cm and 1 Hz, respectively. The squareness of the SW loops (PsSW = PrSW) confirms their complete separation from the respective NSW contributions. The data obtained for the two crystals at other poling field amplitudes and frequencies can be found in the supplementary material. Unlike the conventional ferroelectric hysteresis loops which are throughout continuous, SW, NSW, and FE loops in our investigation show a discontinuity at E = 0, as each of them is measured in two steps by separate half triangular waves of opposite polarities [Fig. 1(b)]. In contrast to PMN-PT, for which the switching saturation polarization (PsSW) is about three times to the non-switching (PsNSW) contribution, Sm:PMN-PT has their values comparable to each other. The distinct behavior of the two crystals is not limited to a high poling field of 15 kV/cm, but it exists throughout the measurement range of 1–15 kV/cm [Fig. 2(c)]. An overall small reduction in the magnitude of ferroelectric polarization of PMN-PT by samarium doping is consistent with that observed in a previous investigation.44. F. Li, M. J. Cabral, B. Xu, Z. Cheng, E. C. Dickey, J. M. LeBeau, J. Wang, J. Luo, S. Taylor, W. Hackenberger, L. Bellaiche, Z. Xu, L.-Q. Chen, T. R. Shrout, and S. Zhang, Science 364(6437), 264 (2019). https://doi.org/10.1126/science.aaw2781Figure 3 compares the evolution of switching loops for PMN-PT and Sm:PMN-PT, respectively, as a function of frequency [Figs. 3(a) and 3(b)] and amplitude [Figs. 3(c) and 3(d)] of the poling field. Evolution of switching loops suggests a gradual increase in the saturation polarization (PSSW) in Sm:PMN-PT with increasing field as compared to PMN-PT, which shows a sharp increase initially, followed by saturation. To perform the dynamic scaling analysis, area bound by the switching loops ⟨ASW⟩ was determined for the two crystals as a function of frequency and amplitude of the poling field.Figures 4(a) and 4(b) compare the variation of ⟨ASW⟩ with the amplitude of the poling field for PMN-PT and Sm:PMN-PT, respectively. Plateauing of ⟨ASW⟩ at E< 15 kV/cm confirms that both the crystals have achieved fairly stable domain configurations at this field. For PMN-PT, sharp inflections in ⟨ASW⟩ vs E curves at ∼ 4 kV/cm suggest that the crystal follows two distinct power laws at low (4 kV/cm) field regimes, respectively. The poor fitting of single power law to the curves obtained for Sm:PMN-PT suggests that it also follows two distinct power laws at low and high fields; however, the inflection point is not so apparent in this case. This phenomenon is not unique to Sm:PMN-PT and PMN-PT crystals as several others (KNN, BaTiO3, etc.) have been reported to follow the distinct power laws at high and low fields, respectively.99. S. Gupta and S. Priya, Appl. Phys. Lett. 98(24), 242906 (2011). https://doi.org/10.1063/1.3600058 The values of the exponent α for the two crystals were determined by fitting ⟨ASW⟩ according to the power law (⟨ASW⟩∝ Eα) in the low (4 kV/cm) field regimes with R2>0.99 and 0.98, respectively. Except for PMN-PT at EFig. 4(c)], in other cases, the parameter α maintains a near-constant behavior with frequency. The average values of α in low and high field regimes are, respectively, 7.87 and 1.15 for PMN-PT in comparison with 3.10 and 1.97 for Sm:PMN-PT.In the power law relationship (⟨ASW⟩∝ Eα), higher value of α implies a more favorable response of the crystallographic domains to the poling field. For PMN-PT, the average values of α as well as Fig. 4(a) suggests a sharp response of crystallographic domains at low fields followed by an abysmal reciprocation at E > 4 kV/cm. On the other hand, Sm:PMN-PT exhibited an intermediate response at low as well as high fields. For both the crystals, poling along ⟨001⟩pc results in the switching of domains between +4R and −4R states involving domain reorientations by 71°, 109°, and 180°, respectively.1818. S. Gupta, Ferroelectric Materials for Energy Harvesting and Storage ( Elsevier, 2020), p. 1; S. Gupta, A. Belianinov, M. Baris Okatan, S. Jesse, S. V. Kalinin, and S. Priya, Appl. Phys. Lett. 104(17), 172902 (2014). https://doi.org/10.1063/1.4874648 Among them, 180° domain switching does not involve any strain energy cost, so is the easiest (in terms of energy required) one to take place. However, due to the zero-strain associated with the process, it only contributes to the polarization, with no contribution to piezoelectricity. On the other hand, 71° and 109° domain reorientations take place at comparatively higher fields and contribute to both polarization and piezoresponse.88. S. Gupta and R. L. Mahajan, Comput. Mater. Sci. 200, 110798 (2021). https://doi.org/10.1016/j.commatsci.2021.110798Our previous theoretical work demonstrated that in the course of achieving the minimum potential energy in a multi-domain system under electric field, ferroelectric (180°) and ferroelastic (71° and 109°) domain switching processes are complementary and mutually competitive to each other.88. S. Gupta and R. L. Mahajan, Comput. Mater. Sci. 200, 110798 (2021). https://doi.org/10.1016/j.commatsci.2021.110798 In PMN-PT, a higher switching of domains by 180° at low fields, understandably, results in their unavailability for ferroelastic switching at high fields. On the contrary, Sm:PMN-PT exhibits a relatively feeble response (in terms of 180° domain switching) at low fields but is able to maintain it at higher fields where ferroelastic switching is most likely. This interpretation of the values of parameter α is not only consistent with the higher piezoelectric activity in Sm:PMN-PT but also explains the higher switching polarization and its sharp rise at low fields in PMN-PT [Figs. 3(c) and 3(d)]. Hence, one can interpret that as compared to PMN-PT, higher ferroelastic switching in Sm:PMN-PT could be one of the possible reasons of higher piezoelectric response of the latter.The exponent of f in power law (β) is a measure of the response time of domains under the varying electric field. From the point of view of functional properties, lower response time is desired. However, under a cyclic electric field, domains do require a finite time to reach the equilibrium state. The response time of domains is higher for the lower amplitude of the poling field and vice versa. If the response time of domains is higher than the inverse of the frequency of the applied field, domains are not able to follow the applied signal, resulting in a poor functional response. This scenario is reflected in a stronger dependence of ⟨ASW⟩ on frequency (higher negative value of parameter β). On the other hand, a frequency independent ⟨ASW⟩ (β=0) suggests that the response time of domains is much smaller than the inverse of f in the measured range. A comparison of the frequency response of the two crystals [Figs. 4(c) and 4(d)] shows better stability of ⟨ASW⟩ in the entire poling field range for Sm:PMN-PT as compared to that for PMN-PT, which has large negative value of β up to 6 kV/cm. On average, the higher negative value of β for PMN-PT (−0.23), as compared to its counterpart for Sm:PMN-PT (−0.12) suggests a lower response time of domains in Sm:PMN-PT.In contrast to the switching loops, which are characteristic of the domain reversal process, non-switching loops (Fig. 5) reveal the dynamics of constitutive ions (within the unit cell) in response to the applied field. The hysteresis in the non-switching loops can be attributed to the time-delayed, frictional response of the lattice in the presence of a varying electric field.19–2119. G. Liu, S. Zhang, W. Jiang, and W. Cao, Mater. Sci. Eng., R 89, 1 (2015). https://doi.org/10.1016/j.mser.2015.01.00220. K.-C. Kao, Dielectric Phenomena in Solids: With Emphasis on Physical Concepts of Electronic Processes ( Academic Press, Amsterdam, Boston, 2004), p. 581.21. D. Damjanovic, The Science of Hysteresis ( Elsevier, 2006), p. 337. The orientation polarization response, which is expected to be high in domain-engineered crystals, is particularly slow and could be the major source of hysteresis in the case of PMN-PT and Sm:PMN-PT.19,2019. G. Liu, S. Zhang, W. Jiang, and W. Cao, Mater. Sci. Eng., R 89, 1 (2015). https://doi.org/10.1016/j.mser.2015.01.00220. K.-C. Kao, Dielectric Phenomena in Solids: With Emphasis on Physical Concepts of Electronic Processes ( Academic Press, Amsterdam, Boston, 2004), p. 581. To reveal any difference in the dynamics of intrinsic polarization between PMN-PT and Sm:PMN-PT crystals, a similar power law scaling (⟨ASW⟩∝ Eδfη) was performed on the non-switching loops as well. As shown in Fig. 5, not only the non-switching loops for Sm:PMN-PT look more saturated than their counterparts for PMN-PT, but also the magnitude of PSNSW for the former is higher than that for the latter. Higher magnitude of polarization in Sm:PMN-PT could be ascribed to its enhanced intrinsic dielectric constant with respect to PMN-PT.3,4,203. F. Li, D. Lin, Z. Chen, Z. Cheng, J. Wang, C. Li, Z. Xu, Q. Huang, X. Liao, L.-Q. Chen, T. R. Shrout, and S. Zhang, Nat. Mater. 17, 349 (2018). https://doi.org/10.1038/s41563-018-0034-44. F. Li, M. J. Cabral, B. Xu, Z. Cheng, E. C. Dickey, J. M. LeBeau, J. Wang, J. Luo, S. Taylor, W. Hackenberger, L. Bellaiche, Z. Xu, L.-Q. Chen, T. R. Shrout, and S. Zhang, Science 364(6437), 264 (2019). https://doi.org/10.1126/science.aaw278120. K.-C. Kao, Dielectric Phenomena in Solids: With Emphasis on Physical Concepts of Electronic Processes ( Academic Press, Amsterdam, Boston, 2004), p. 581.Figure 6 compares the area ⟨ANSW⟩ enclosed by the non-switching loops for PMN-PT and Sm:PMN-PT as a function of frequency [Figs. 6(a) and 6(b)] and amplitude [Figs. 6(c) and 6(d)] of the applied field. In spite of higher PSNSW [Fig. 1(d)] for Sm:PMN-PT, comparable values of ⟨ANSW⟩ for the two crystals at low frequencies suggest slimmer loops for Sm:PMN-PT and hence the ease of polarization reversal (in terms of coercivity) as compared to that for PMN-PT. Considering the evolution of ⟨ANSW⟩ with E, PMN-PT exhibited a frequency-dependent [Fig. 6(a)] growth according to the power law (⟨ANSW⟩∝ Eδ) with average value of δ being 1.88. On the other hand, Sm:PMN-PT exhibits a frequency-independent [Fig. 6(b)], near-linear growth of ANSW with an average value of δ being 1.2. The exponent of E is a measure of the ability of polarization (arising due to the relative movement of constitutive ions in this case) to respond to the varying amplitude of the electric field. A higher value of δ for PMN-PT suggests a more favorable response from the non-switching polarization for it as compared to the Sm doped counterpart. More importantly, unlike the case of switching loops, ⟨ANSW⟩ vs E curves followed the same power law in the entire range of 1–15 kV/cm with R2>0.98. This observation suggests that a single mechanism of intrinsic contribution spans over the measured electric field range. Any sharp change in ⟨ANSW⟩ vs E curves or change in power law would have suggested a polarization rotation, which is not the case here. Comparing the variation of ⟨ANSW⟩ with frequency for the two crystals, Sm:PMN-PT exhibits a near-constant behavior of η in the range of 1–200 Hz (an average value of −0.02) as compared to a frequency-dependent behavior (average value of η=−0.11) in PMN-PT. Again, a relatively low frequency dependence of exponent η suggests a smaller response time of non-switching polarization in Sm:PMN-PT, which is an advantageous factor as far as the functional properties of Sm:PMN-PT are concerned.

In summary, the areas bound by the switching and non-switching loops for ⟨001⟩pc oriented PMN-PT and Sm:PMN-PT crystals were subjected to dynamic scaling analysis according to the power law relations. The switching loops for PMN-PT were found to obey the power laws ⟨A⟩SW ∝ E7.87f−0.23 and ⟨A⟩SW ∝ E1.15f−0.23 at low (E<4 kV/cm) and high (E>4 kV/cm) fields, respectively. On the other hand, its samarium doped counterpart has been found to follow the relations ⟨A⟩SW ∝ E3.1f−0.12 (E<4 kV/cm) and ⟨A⟩SW ∝ E1.97f−0.12 (E>4 kV/cm). For the switching loops, the relatively higher value of E-exponent for PMN-PT at low fields and for Sm:PMN-PT at high fields suggest that the domain reorientation process in the former is dominated by ferroelectric switching, whereas the latter is more apt for ferroelastic switching. Non-switching loops for the two crystals follow the power law relations ⟨A⟩SW ∝ E1.88f−0.11 and ⟨A⟩SW ∝ E1.2f−0.02, respectively. The smaller negative value of f-exponent for Sm:PMN-PT for switching as well non-switching loops suggest a small response time of polarizations in both the cases, which is again consistent with the high functional response in Sm:PMN-PT crystal. In addition to the crystallographic and thermodynamic approaches adopted earlier, our work adds another facet to the understanding of the mechanism responsible for the high piezoelectric activity in the samarium doped PMN-0.28PT crystal. As the mechanism of high piezoresponse in samarium doped PMN-xPT ceramics is likely to be the same, our findings are equally valid for them as well.

See the supplementary material for the ferroelectric, switching, and non-switching loops obtained for PMN-PT and Sm:PMN-PT crystals at different electric fields and frequencies.

The authors gratefully acknowledge the financial support from VT India.

Conflict of Interest

The authors have no conflicts to disclose.

Author Contributions

Shashaank Gupta: Conceptualization (lead); Data curation (lead); Formal analysis (lead); Writing – original draft (lead); Writing – review & editing (lead). Myoor K. Padmanabhan: Supervision (supporting). Roop Mahajan: Supervision (supporting); Writing – review & editing (supporting).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

REFERENCES

1. S.-E. Park and T. R. Shrout, J. Appl. Phys. 82(4), 1804 (1997). https://doi.org/10.1063/1.365983, Google ScholarScitation, ISI2. B. Noheda, D. E. Cox, G. Shirane, J. Gao, and Z. G. Ye, Phys. Rev. B 66(5), 054104 (2002); https://doi.org/10.1103/PhysRevB.66.054104, Google ScholarCrossref
B. Noheda, Curr. Opin. Solid State Mater. Sci. 6(1), 27 (2002). https://doi.org/10.1016/S1359-0286(02)00015-3, , Google ScholarCrossref3. F. Li, D. Lin, Z. Chen, Z. Cheng, J. Wang, C. Li, Z. Xu, Q. Huang, X. Liao, L.-Q. Chen, T. R. Shrout, and S. Zhang, Nat. Mater. 17, 349 (2018). https://doi.org/10.1038/s41563-018-0034-4, Google ScholarCrossref, ISI4. F. Li, M. J. Cabral, B. Xu, Z. Cheng, E. C. Dickey, J. M. LeBeau, J. Wang, J. Luo, S. Taylor, W. Hackenberger, L. Bellaiche, Z. Xu, L.-Q. Chen, T. R. Shrout, and S. Zhang, Science 364(6437), 264 (2019). https://doi.org/10.1126/science.aaw2781, Google ScholarCrossref5. C. Zhao, F. Li, S. Zhang, S. Li, and J. L. Jones, J. Appl. Phys. 126(7), 075101 (2019). https://doi.org/10.1063/1.5089477, Google ScholarScitation, ISI6. M. Davis, M. Budimir, D. Damjanovic, and N. Setter, J. Appl. Phys. 101(5), 054112 (2007). https://doi.org/10.1063/1.2653925, Google ScholarScitation, ISI7. A. A. Levin, A. I. Pommrich, T. Weißbach, P. Paufler, and D. C. Meyer, Int. J. Mater. Res. 98(8), 700 (2007); https://doi.org/10.3139/146.101521,

留言 (0)

沒有登入
gif