Next Article in Journal
The A, B and C’s of Silicone Breast Implants: Anaplastic Large Cell Lymphoma, Biofilm and Capsular Contracture
Next Article in Special Issue
Domain Diversity and Polarization Switching in Amino Acid β-Glycine
Previous Article in Journal
Polyvinyl Alcohol Reinforced Flame-Retardant Polyacrylonitrile Composite Fiber Prepared by Boric Acid Cross-Linking and Phosphorylation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis, Giant Dielectric, and Pyroelectric Response of [001]-Oriented Pr3+ Doped Pb(Mg1/3Nb2/3)O3-PbTiO3 Ferroelectric Nano-Films Grown on Si Substrates

1
Thin Film and Optical Manufacturing Technology, Key Laboratory of Ministry of Education, Xi’an Technological University, Xi’an 710032, China
2
Department of Basic Science, Air Force Engineering University, Xi’an 710051, China
*
Author to whom correspondence should be addressed.
Materials 2018, 11(12), 2392; https://doi.org/10.3390/ma11122392
Submission received: 27 October 2018 / Revised: 20 November 2018 / Accepted: 25 November 2018 / Published: 28 November 2018
(This article belongs to the Special Issue Nanoscale Ferroelectrics and Their Applications)

Abstract

:
The [001]-oriented Pr3+ doped Pb(Mg1/3Nb2/3)O3-0.30PbTiO3 (Pr-PMN-PT) thin films with a composition near the morphotropic phase boundary (MPB) were synthesized by a sol–gel method. The crystal structure was characterized using X-ray diffraction. It was found that a single perovskite phase was achieved in Pr-PMN-PT thin films annealed at 650 °C for 3 min. The dielectric constant (εr) value was 2400 in 2.5% Pr-PMN-PT thin films at room temperature, 110% higher than that of pure PMN-PT samples. Through 2.5% Pr3+ doping, remanent polarization (Pr) and coercive field (Ec) values increased from 11.5 μC/cm2 and 35 kV/cm to 17.3 μC/cm2 and 63.5 kV/cm, respectively, in PMN-PT thin films. The leakage current densities of pure and 2.5% Pr-PMN-PT thin films were on the order of 1.24 × 10−4 A/cm2 and 5.8 × 10−5 A/cm2, respectively, at 100 kV/cm. A high pyroelectric coefficient (py) with a value of 167 μC/m2K was obtained in 2.5% Pr-PMN-PT thin films on Si substrate, which makes this material suitable for application in infrared detectors.

Graphical Abstract

1. Introduction

Pb(Mg1/3Nb2/3)O3-PbTiO3 (PMN-PT) bulk piezoelectric materials with compositions near the morphotropic phase boundary have attracted much attention as a candidate piezoelectric material for device application in the fields of microelectromechanical system (MEMS) and nanoelectromechanical systems (NEMS) [1,2,3]—such as infrared (IR) detectors [4], cantilever-type piezoelectric devices [5,6], multilayer capacitors [7], and ultrasonic transducers [8,9,10]—due to the high field strain (1.7%) [11], pyroelectric coefficient (~1500 μC/m2K) [12], and longitudinal coupling factor (~95%) [13,14]. Especially for uncooled IR detectors, pyroelectricity is indeed one of the most promising principles. Compared with bulk crystals, thin films have the advantage of low-cost, reduced material consumption, small cell size, energy efficiency, and ease of integration in Si semiconductor technology. For PMN-based bulk crystals, the pure perovskite structure can easily be synthesized using the two-step columbite precursor method [15,16,17]. On the contrary, the second pyrochlore phase in PMN-PT thin films is related to many factors, for example annealing temperature/time [18], sample thickness [19], excess Pb content [20,21], etc., which restrict the pyroelectric, piezoelectric, and dielectric properties in PMN-PT thin films.
Many methods and approaches—such as radio frequency (RF) magnetron sputtering [22], pulsed laser deposition (PLD) [4], chemical vapor deposition (CVD) [23], chemical solution deposition (CSD) [24], and so forth—have been adopted to fabricate high-quality PMN-PT thin films. Sol–gel processing, a kind of chemical solution deposition, has many advantages that make it suitable for obtaining high-quality PMN-PT thin films on Si substrate for industrial applications [25,26]. Dielectric constants and remanent polarization in slim P-E hysteresis loops were on the order of 1692 and 13.31 μC/cm2 respectively in the PMN-0.30PT thin films using sol-gel technology [27]. Recently, a giant pyroelectric response (−550 μCm−2K−1) driven by the electric field was obtained in 0.68Pb(Mg1/3Nb2/3)O3-0.32PbTiO3/Ba0.5Sr0.5RuO3 hetero structures grown on [110]-oriented NdScO3 single crystals by pulsed-laser deposition (PLD) [4] indicating good potential for low dimensional pyroelectric devices.
Rare earth elements such as Er3+, La3+, and so forth have been used to improve the up-conversion photoluminescence in bulk piezoelectric materials [28,29]. Recently, an ultrahigh piezoelectric coefficient (d33 = 1500 pC/N) was obtained in Sm-doped PMN-PT polycrystalline [30] based on polar nano-region (PNR) design in relaxor based ferroelectric materials. For Pb(Zr,Ti)O3 (PZT) thin films, the dielectric and piezoelectric properties were improved by La doping [31]. Until now, little research has focused on the pyroelectric or piezoelectric properties of rare earth element-doped relaxor-PT (PMN-PT or PIN-PMN-PT) thin films. With the aim of low cost and high performance materials for MEMS device fabrication, the Pr3+ doped PMN-PT thin films were prepared by the sol–gel method at various annealed temperatures and times. The phase structure of the samples was analyzed by X-ray diffraction. Then, the surface morphology of Pr-PMN-PT thin films was measured by scanning electron microscopy, atomic force microscopy, and piezoresponse force microscopy. Finally, dielectric, pyroelectric, and ferroelectric properties of Pr-PMN-PT thin films were investigated.

2. Materials and Methods

Before sol–gel processing, a TiO2 diffusion barrier layer was first fabricated by oxidizing the Ti metal layer as follows: Ti films of 60 nm thickness were deposited on SiO2/Si (100) substrate by DC magnetron sputtering in an Ar gas environment at room temperature. The sputtering power was 175 W. The base vacuum of Ti sputtering was higher than 8 × 10−4 Pa. The ratio of Ar:O2 was 90:10. After deposition, the Ti layer was annealed in a rapid thermal processing (RTP) furnace at 700 °C for 1 h. Then, a Pt bottom electrode of 80 nm thickness was deposited on the TiO2 diffusion barrier layer. The sputtering parameters were the same as the Ti metal layer deposition. The sheet resistance was about 2.85 Ω/sq, which was measured by the four probe method.
After that, Pr3+ doped Pb(M1/3Nb2/3)O3-0.30PbTiO3 (xPr-(1 − x)(PMN-0.30PT), x = 1%, 2%, 2.5%, and 3%) precursor solutions were prepared by the sol–gel method. The lead acetate trihydrate (Pb(CH3COO)2·3H2O), magnesium acetate tetrahydrate (Mg(CH3COO)2·4H2O), niobium ethoxide (Nb(OCH3CH2)5), tetrabutyl titanate (Ti(OC4H9)4), and praseodymium acetyl acetone (C15H21PrO6·xH2O) were used as starting materials to synthesize the Pr-PMN-PT sol–gel solution. An additional 20 mol% of lead acetate trihydrate was added to equalize the Pb volatilization and 5 mol% more magnesium acetate tetrahydrate was mixed to boost the production of perovskite phase. Next, 2-methoxyethanol (C5H8O2) and glacial acetic acid (CH3COOH) were added as solvent and catalyst respectively. Acetyl acetone (C5H8O2) was added as a stabilizer in the Pb-Pr-Mg-Nb-Ti solution. The concentration of the Pr-PMN-PT solution was attenuated to 0.4 mol/L. Then, the Pr-PMN-PT thin films were deposited by spin coating on Pt/TiO2/SiO2/Si substrates at 2000 rpm for 9 s before the spin coating was subsequently sped up to 4000 rpm for 20 s. To evaporate the solvent in solution, the sample was dried at 220 °C for 3 min after each deposition and then pyrolyzed at 450 °C for 5 min. Finally, Pr-PMN-PT films were annealed at 600, 650, and 700 °C for 1–5 min in air by RTP at ramp rates +20 °C/s and −5 °C/s respectively. In order to get 500 nm thick film, this entire process was repeated 10 times. For further electric measurement, Au top electrodes were deposited on Pr-PMN-PT samples using magnetron sputtering through a shadow mask with a 1 mm diameter.
The crystalline structure of Pr-PMN-PT thin films was analyzed by X-ray diffraction (XRD) (Shimadzu X-6000, Kyoto, Japan) with Cu-Kα radiation in the 2theta range of 20°–60° and a scan step length of 0.02 at room temperature. The surface morphologies of the samples were characterized by a field emission scanning electron microscope (FE-SEM) (ZEISS Gemini SEM 500, Jena, Germany), atomic force microscope (AFM) (Bruker Multimode 8, Billerica, MA, USA), and piezoresponse force microscopy (PFM) (Bruker dimension icon, Billerica, MA, USA). For electric measurement, the samples were all of a pure perovskite phase structure due to the high electric performance. The dielectric properties were measured using an impedance analyzer (HP 4284, Hewlett Packard, Palo Alto, CA, USA) with a precision LCR meter connected to a heating stage (Linkam THMSE 600, Waterfield, UK). The P-E hysteresis loops and leakage current were examined by a standard ferroelectric system (aixACCT TF-2000, Aachen, Germany). The pyroelectric coefficient (py) was calculated by the equation as below:
p y ( T ) = I A d T d t   ( μ C / m 2 K ) ,
where I is the pyroelectric current, A is the surface area of the samples, and dT/dt is the rate of temperature cooling.

3. Results and Discussion

3.1. Phase Structure

Avoiding the formation of a pyrochlore phase is a key factor for the synthesis of PMN-PT thin films as the addition of this phase will reduce the piezoelectric and ferroelectric properties of the resulting thin films. In order to restrict the formation of a pyrochlore phase, the annealing conditions for thin film preparation were first investigated. Figure 1a,b show the X-ray diffraction patterns of 2.5% Pr3+ doped PMN-0.30PT thin films deposited on Pt/TiO2/SiO2/Si substrates annealed at 600 °C, 650 °C, and 700 °C for 2 min. The perovskite phase peaks were observed at 2θ = 22.1°, 31.4°, 38.6°, 45.2°, 50.9°, and 56.0°, corresponding to the (001), (110), (111), (002), (220), and (112) reflections respectively. The pyrochlore phase peak (222) at 2θ = 29.4° was clear for samples annealed at 600 °C and 700 °C and was identified as Pb2Nb2O7 by the comparison to JCPDS No. 40-828, which showed the other characteristic diffraction peak at 2θ = 34.1°. Compared to the typical PMN-PT solid solution, the single perovskite structure was obtained at 650 °C, which showed two strong peaks at 2θ = 22.2° and 45.3°, indicating a mainly preferred crystal growth along the (001)c direction. Figure 1c,d present the phase structure of 2.5% Pr-PMN-PT thin films deposited on Pt/TiO2/SiO2/Si substrates annealed at 650 °C for 1 min, 3 min, and 5 min respectively. It was found that 2.5% Pr-PMN-PT thin films were mostly in the perovskite phase with a little pyrochlore phase when annealed for 1 min, which was further crystallized to a single perovskite phase structure after one more minute annealing. However, a peak of the pyrochlore phase was found again in the sample annealed for 5 min, which can be ascribed to the greater lead loss caused by the longer time annealing [19].
The degree of (001) texture in 2.5% Pr-PMN-PT thin film was evaluated by preferred orientation parameter αhkl, which can be calculated by the following formula [32]:
α hkl = I hkl I hkl ,
where Ihkl is the relative intensity of the corresponding diffraction peaks (hkl). As shown in Figure 1a,c, α001 values are 86.6% and 73.6% in Pr-PMN-PT thin films annealed at 650 °C for 2 and 3 min respectively, presenting the highly (001)-oriented textured structure. The preferred orientation was related to many factors, such as substrate materials [33], growth method, doping elements [31], annealing temperatures [34], and so forth. For Pr-PMN-PT thin films, the nucleation was in random direction at the start of thin film growth on the surface of Pt. With further temperature increases up to 650 °C, the (001) direction became the preferential growth direction.

3.2. Surface Morphology and Electric Properties at Room Temperature

Figure 2a,b shows the FE-SEM images of surface and cross-sectional micrographs of Pr-PMN-PT thin films. From these images, it can be seen that the thin films have a dense structure without any cracks. It can further be seen that the quite uniform Pr-PMN-PT, Pt, and TiO2 films are around 450 nm, 80 nm, and 80 nm in thickness respectively, which are in agreement with the values measured by a surface profilometer. Figure 2c presents surface grain size distributions calculated from the image in Figure 2a, showing an average grain size of 100 nm. A typical out-of-plane image with a 3 × 3 μm2 scan area at the surface of 2.5% Pr-PMN-PT thin films is shown in Figure 2d. The grain size is about 100 nm, in accordance with the FE-SEM observation. However, nano-domains could not be found in the picture, which can be analyzed by macroscale level ferroelectric properties, similar to the results in Ref. [35,36]. The 3D AFM image of 2.5% Pr-PMN-PT thin films is shown in Figure 2e. The root mean square roughness (Rrms) is about 2.59 nm with a scan area of 2 × 2 μm2, indicating low interference of electrode contact in electronic measurement. A schematic of the device for further electric investigation is shown in Figure 2f. Two silver wires for electric measurement were bonded to the surface of the Au electrodes.
The dielectric, pyroelectric, and ferroelectric properties in 2.5% Pr-PMN-PT thin films at room temperature are summarized in Table 1. The values of εr, py, and pr in 2.5% Pr-PMN-PT thin films were on the order of 2400, 167 μC/m2K, and 17.3 μC/cm2, showing 110%, 123%, and 50% higher factors respectively than those in pure PMN-PT thin films. Compared with PMN-0.30PT grown by a chemical solution deposition (CSD) method [19], the εr in PMN-0.30PT thin films on Si substrate was lower than that of the sample synthesized on perovskite structure single crystal substrates. A similar tendency of pyroelectric property was found in PMN-0.30PT thin film grown on various substrates. The pr value in 2.5% Pr-PMN-PT thin films was still higher than that of PMN-0.32PT thin films deposited on Ba0.5Sr0.5RuO3/NdScO3 and La0.5Sr0.5CoO3/LaAlO3 crystals using pulsed laser deposition. In contrast, the py value showed an opposite trend, which is attributed to the lattice misfit. Compared with PVDF and normal ferroelectric material, the relaxor-PT thin films showed higher dielectric and pyroelectric properties. In contrast, the response time of the PMN-0.30PT pyroelectric sensor was longer than that of PVDF [37] and LiTaO3 [38].
In order to evaluate the Pr3+ doping effect in PMN-PT thin films, the elements distribution was measured, as shown in Figure 3, corresponding to the surface micrograph presented in Figure 2a. The Pb, Mg, O, Pr, Nb, Ti, Si, and Pt maps were achieved by performing standardless quantification of the energy dispersive X-ray spectrometry (EDS) signals in the Lα1 and Kα1 peaks. From these pictures, it can be seen that the Pb, Mg, Nb, Ti, Si, and Pt elements were homogeneously distributed in the local area. It is interesting to note that the Pr element is mostly distributed along the grain boundaries of thin films, compared to Figure 2a.

3.3. Ferroelectric Properties

Figure 4a presents the ferroelectric properties of Pr-PMN-PT thin films with different Pr3+ ratios of 0, 1.5%, and 2.5%. The addition of Pr3+ in PMN-PT thin films increased the value of remanent polarization (Pr) and coercive field (Ec) from 11.5 μC/cm2 and 35 kV/cm to 17.3 μC/cm2 and 63.5 kV/cm respectively. The leakage current densities of pure and 2.5% Pr-PMN-PT thin films were approximately 1.24 × 10−4 A/cm2 and 5.8 × 10−5 A/cm2 respectively at 100 kV/cm, as shown in Figure 4b. A donor dopant in PMN-PT samples was poled at the higher electric field due to the higher Ec in Pr-PMN-PT thin films, indicating that the domain structure was easily reoriented owing to the presence of lead vacancies near the grain boundaries [44,45], corresponding to the Pr3+ ion distribution shown in Figure 3. Those effects also led to a higher dielectric constant of Pr3+ doped compositions compared to the undoped counterpart, as shown in Table 1.

3.4. Temperature Dependent of Dielectric and Pyroelectric Properties

Figure 5a shows dielectric constant (εr) and dielectric loss (tanδ) of Pr-PMN-PT thin films in the frequency range of 20 Hz to 200 kHz at room temperature. We see that the εr value decreased slightly with an increasing frequency, which is induced by the polarization decrease due to charge accumulation at low frequency. Meanwhile, the dielectric loss showed a very quick increase in the range 104–2 × 105 Hz, indicating conduction through the electrode surface and grain boundary capacitance interference in poled samples at high frequency. Figure 5b presents the dielectric properties dependent on temperature for Pr-PMN-PT thin film at 100 Hz and 1 kHz. The εr curve presents a broad peak centered at around 100 °C, which corresponds to the diffuse ferroelectric to paraelectric phase transition [10,17]. The value of dielectric loss values is about 1.8% in the temperature range from 25 to 100 °C, which indicates good electrode bonding for further device integrating.
Figure 6a shows the pyroelectric coefficient (py) measured at room temperature for the obtained PMN-PT thin films as a function of Pr3+ doping content. The py in the thin films with 2.5% dopant presented the maximum value on the order of 167 μC/m2K, which is almost a two-fold increase over that of pure PMN-PT thin films, offering a new way for enhancing the pyroelectric performance in relaxor-PT thin films for use in infrared detectors. The temperature dependence of py values is illustrated in Figure 6b. In the temperature range of −50–100 °C, the pyroelectric coefficient showed a slight fluctuation (~10%), indicating good thermal stability for application in infrared detectors integrated into Si semiconductor devices.

4. Conclusions

The pure and Pr3+ doped PMN-PT ferroelectric thin films with a single perovskite structure were prepared by a sol–gel processing and investigated by X-ray diffraction. The microstructure of Pr-PMN-PT samples was characterized by scanning electron microscopy, atomic force microscopy, and piezoresponse force microscopy. The average grain size was about 100 nm. It was found that Pr3+ ions were located at the edge of grain boundaries measured by energy dispersive X-ray spectroscopy. For 2.5% Pr3+ doping, the dielectric constant, pyroelectric coefficient, and remenant polarization all reached a maximum. Importantly, the dielectric constant in PMN-PT thin films increased from 1140 to 2400 through 2.5% Pr3+ doping. Such enhancement was ascribed to the domain wall motion induced by the Pr3+ ion distribution around the grain boundary. Finally, this rare earth dopant presents a useful method for developing high performance relaxor-PT thin films for use in energy transducers and photodetectors.

Author Contributions

C.C., D.Z., Y.S., X.S., and D.L. synthesized the films and electric measurement; D.Z., J.W., and Y.S. performed the structure experiments; C.C., W.L., S.Z., and D.L. analyzed the data; C.C. and D.L. wrote the manuscript and all authors discussed the results; the project was conceived and designed by D.L.

Funding

This work was supported by the support from National Natural Science Foundation of China (No. 51502232), National Basic Research Project (JCKY2016208A002), Advanced Manufacturing Project (41423020111).

Acknowledgments

C.C. wants to thank the support from Xi’an Key Laboratory of Intelligent Detection and Perception (201805061ZD12CG45), Shaanxi provincial industry innovation chain project (2018ZDCXL-GY-08-02-01).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Baek, S.-H.; Rzchowski, M.S.; Aksyuk, V.A. Giant piezoelectricity in PMN-PT thin films: Beyond PZT. MRS Bull. 2012, 37, 1022–1029. [Google Scholar] [CrossRef]
  2. Cordero, F. Elastic Properties and Enhanced Piezoelectric Response at Morphotropic Phase Boundaries. Materials 2015, 8, 5452. [Google Scholar] [CrossRef] [PubMed]
  3. Lin, D.; Li, Z.; Cheng, Z.Y.; Xu, Z.; Yao, X. Electric-field-induced polarization fatigue of [001]-oriented Pb(Mg1/3Nb2/3)O3–0.32PbTiO3 single crystals. Solid State Commun. 2011, 151, 1188–1191. [Google Scholar] [CrossRef]
  4. Pandya, S.; Wilbur, J.; Kim, J.; Gao, R.; Dasgupta, A.; Dames, C.; Martin, L.W. Pyroelectric energy conversion with large energy and power density in relaxor ferroelectric thin films. Nat. Mater. 2018, 17, 432. [Google Scholar] [CrossRef] [PubMed]
  5. Lee, S.H.; Jeong, C.K.; Hwang, G.-T.; Lee, K.J. Self-powered flexible inorganic electronic system. Nano Energy 2015, 14, 111–125. [Google Scholar] [CrossRef]
  6. Zhang, S.; Li, F.; Lee, H.J.; Lin, D.; Shrout, T.R.; Luo, J.; Meyer, R.J. Evaluation of PMN-PT based crystals for various applications. In Proceedings of the 2011 International Symposium on Applications of Ferroelectrics (ISAF/PFM) and 2011 International Symposium on Piezoresponse Force Microscopy and Nanoscale Phenomena in Polar Materials, Vancouver, BC, Canada, 24–27 July 2011; pp. 1–4. [Google Scholar]
  7. Moya, X.; Defay, E.; Mathur, N.D.; Hirose, S. Electrocaloric effects in multilayer capacitors for cooling applications. MRS Bull. 2018, 43, 291–294. [Google Scholar] [CrossRef]
  8. Zhang, T.; Ou-Yang, J.; Yang, X.; Zhu, B. Transferred PMN-PT Thick Film on Conductive Silver Epoxy. Materials 2018, 11, 1621. [Google Scholar] [CrossRef] [PubMed]
  9. Chen, Y.; Lam, K.-H.; Zhou, D.; Yue, Q.; Yu, Y.; Wu, J.; Qiu, W.; Sun, L.; Zhang, C.; Luo, H.; et al. High Performance Relaxor-Based Ferroelectric Single Crystals for Ultrasonic Transducer Applications. Sensors 2014, 14, 13730. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Lin, D.; Lee, H.J.; Zhang, S.; Li, F.; Li, Z.; Xu, Z.; Shrout, T.R. Influence of domain size on the scaling effects in Pb(Mg1/3Nb2/3)O3–PbTiO3 ferroelectric crystals. Scripta Mater. 2011, 64, 1149–1151. [Google Scholar] [CrossRef] [PubMed]
  11. Park, S.-E.; Shrout, T.R. Ultrahigh strain and piezoelectric behavior in relaxor based ferroelectric single crystals. J. Appl. Phys. 1997, 82, 1804–1811. [Google Scholar] [CrossRef]
  12. Choi, S.W.; Shrout, R.T.R.; Jang, S.J.; Bhalla, A.S. Dielectric and pyroelectric properties in the Pb(Mg1/3Nb2/3)O3-PbTiO3 system. Ferroelectrics 1989, 100, 29–38. [Google Scholar] [CrossRef]
  13. Li, F.; Zhang, S.; Xu, Z.; Chen, L.-Q. The Contributions of Polar Nanoregions to the Dielectric and Piezoelectric Responses in Domain-Engineered Relaxor-PbTiO3 Crystals. Adv. Funct. Mater. 2017, 27, 1700310. [Google Scholar] [CrossRef]
  14. Kazys, R.; Sliteris, R.; Sestoke, J. Air-Coupled Ultrasonic Receivers with High Electromechanical Coupling PMN-32%PT Strip-Like Piezoelectric Elements. Sensors 2017, 17, 2365. [Google Scholar] [CrossRef] [PubMed]
  15. Swartz, S.L.; Shrout, T.R. Fabrication of perovskite lead magnesium niobate. Mater. Res. Bull. 1982, 17, 1245–1250. [Google Scholar] [CrossRef]
  16. Lin, D.; Li, Z.; Li, F.; Xu, Z.; Yao, X. Characterization and piezoelectric thermal stability of PIN–PMN–PT ternary ceramics near the morphotropic phase boundary. J. Alloys Compd. 2010, 489, 115–118. [Google Scholar] [CrossRef]
  17. Lin, D.; Li, Z.; Zhang, S.; Xu, Z.; Yao, X. Electric-field and temperature induced phase transitions in Pb(Mg1/3Nb2/3)O3–0.3PbTiO3 single crystals. J. Appl. Phys. 2010, 108, 034112. [Google Scholar] [CrossRef]
  18. Shen, B.; Wang, J.; Pan, H.; Chen, J.; Wu, J.; Chen, M.; Zhao, R.; Zhu, K.; Qiu, J. Effects of annealing process and the additive on the electrical properties of chemical solution deposition derived 0.65Pb(Mg1/3Nb2/3)O3–0.35PbTiO3 thin films. J. Mater. Sci. Mater. Electron. 2018, 29, 16997–17002. [Google Scholar] [CrossRef]
  19. Keech, R.; Morandi, C.; Wallace, M.; Esteves, G.; Denis, L.; Guerrier, J.; Johnson-Wilke, R.L.; Fancher, C.M.; Jones, J.L.; Trolier-McKinstry, S. Thickness-dependent domain wall reorientation in 70/30 lead magnesium niobate-lead titanate thin films. J. Am. Ceram. Soc. 2017, 100, 3961–3972. [Google Scholar] [CrossRef]
  20. Keech, R.; Shetty, S.; Wang, K.; Trolier-McKinstry, S. Management of Lead Content for Growth of {001}-Oriented Lead Magnesium Niobate-Lead Titanate Thin Films. J. Am. Ceram. Soc. 2016, 99, 1144–1146. [Google Scholar] [CrossRef]
  21. Borman, T.M.; Zhu, W.; Wang, K.; Ko, S.W.; Mardilovich, P.; Trolier-McKinstry, S.E. Effect of lead content on the performance of niobium-doped {100} textured lead zirconate titanate films. J. Am. Ceram. Soc. 2017, 100, 3558–3567. [Google Scholar] [CrossRef]
  22. Herdier, R.; Detalle, M.; Jenkins, D.; Soyer, C.; Remiens, D. Piezoelectric thin films for MEMS applications—A comparative study of PZT, 0.7PMN-0.3PT and 0.9PMN-0.1PT thin films grown on Si by r.f. magnetron sputtering. Sens. Actuators A Phys. 2008, 148, 122–128. [Google Scholar] [CrossRef]
  23. Yutaka, T.; Kosuke, S.; Hiroshi, T.; Kunisaburo, T. Preparation of Lead Magnesium Niobate Titanate Thin Films by Chemical Vapor Deposition. Jpn. J. Appl. Phys. 1995, 34, 5083. [Google Scholar]
  24. Keech, R.; Shetty, S.; Kuroda, M.A.; Liu, X.H.; Martyna, G.J.; Newns, D.M.; Trolier-McKinstry, S. Lateral scaling of Pb(Mg1/3Nb2/3)O3-PbTiO3 thin films for piezoelectric logic applications. J. Appl. Phys. 2014, 115, 234106. [Google Scholar] [CrossRef]
  25. Kim, Y.-H.; Heo, J.-S.; Kim, T.-H.; Park, S.; Yoon, M.-H.; Kim, J.; Oh, M.S.; Yi, G.-R.; Noh, Y.-Y.; Park, S.K. Flexible metal-oxide devices made by room-temperature photochemical activation of sol–gel films. Nature 2012, 489, 128. [Google Scholar] [CrossRef] [PubMed]
  26. Karataş, Ş.; El-Nasser, H.M.; Al-Ghamdi, A.A.; Yakuphanoglu, F. High Photoresponsivity Ru-doped ZnO/p-Si Heterojunction Diodes by the Sol-gel Method. Silicon 2018, 10, 651–658. [Google Scholar] [CrossRef]
  27. Zhou, D.; Sun, H.; Liu, X.; Sui, H.; Guo, Q.; Liu, P.; Ruan, Y. Effects of annealing temperature on structure and electrical properties of sol-gel derived 0.65PMN-0.35PT thin film. Ceram. Int. 2017, 43, 5901–5906. [Google Scholar] [CrossRef]
  28. Yao, Y.; Luo, L.; Li, W.; Zhou, J.; Wang, F. An intuitive method to probe phase structure by upconversion photoluminescence of Er3+ doped in ferroelectric Pb(Mg1/3Nb2/3)O3-PbTiO3. Appl. Phys. Lett. 2015, 106, 082906. [Google Scholar] [CrossRef]
  29. Yang, M.; Long, S.; Yang, X.; Lin, S.; Zhu, Y.; Ma, D.; Wang, B. Temperature-Dependent and Threshold Behavior of Sm3+ Ions on Fluorescence Properties of Lithium Niobate Single Crystals. Materials 2018, 11, 2058. [Google Scholar] [CrossRef] [PubMed]
  30. Li, F.; Lin, D.; Chen, Z.; Cheng, Z.; Wang, J.; Li, C.; Xu, Z.; Huang, Q.; Liao, X.; Chen, L.-Q.; et al. Ultrahigh piezoelectricity in ferroelectric ceramics by design. Nat. Mater. 2018, 17, 349. [Google Scholar] [CrossRef] [PubMed]
  31. Yang, L.; Zheng, D.-Y.; Guo, K.-X.; Zhao, W.-N.; Peng, Z.-H.; Peng, G.-G.; Zhou, T. Mechanism of grain growth and excellent polarization, dielectric relaxtion of La3+, Nd3+ modified PZT nano-films prepared by sol–gel technique. J. Mater. Sci. Mater. Electron. 2018, 29, 18011–18019. [Google Scholar] [CrossRef]
  32. Feng, M.; Wang, W.; Ke, H.; Rao, J.C.; Zhou, Y. Highly (111)-oriented and pyrochlore-free PMN–PT thin films derived from a modified sol–gel process. J. Alloys Compd. 2010, 495, 154–157. [Google Scholar] [CrossRef]
  33. Gabor, U.; Spreitzer, M.; Uršič, H.; Tchernychova, E.; Samardžija, Z.; Wu, W.J.; Suvorov, D. Structural peculiarities of 0.67Pb(Mg1/3Nb2/3)O3–0.33PbTiO3 thin films grown directly on SrTiO3 substrates. J. Eur. Ceram. Soc. 2018, 38, 4453–4462. [Google Scholar] [CrossRef]
  34. Batra, V.; Kotru, S. Temperature-controlled-orientation dependence of electrical properties of Pb0.95La0.05Zr0.46Ti0.54O3 thin films. J. Vac. Sci. Technol. B 2018, 36, 052908. [Google Scholar] [CrossRef]
  35. Detalle, M.; Ferri, A.; Da Costa, A.; Desfeux, R.; Soyer, C.; Rémiens, D. Nanoscale study by piezoresponse force microscopy of relaxor 0.7Pb(Mg1/3Nb2/3)O3–0.3PbTiO3 and 0.9Pb(Mg1/3Nb2/3)O3–0.1PbTiO3 thin films grown on platinum and LaNiO3 electrodes. Thin Solid Films 2010, 518, 4670–4674. [Google Scholar] [CrossRef]
  36. Saranya, D.; John, N.S.; Parui, J.; Krupanidhi, S.B. Studies on Field Dependent Domain Structures in Multi-Grained 0.85PbMg1/3Nb2/3O3–0.15PbTiO3 Thin Films by Scanning Force Microscopy. Integr. Ferroelectr. 2012, 134, 39–47. [Google Scholar] [CrossRef]
  37. Pullano, S.; Mahbub, I.; Islam, S.; Fiorillo, A. PVDF Sensor Stimulated by Infrared Radiation for Temperature Monitoring in Microfluidic Devices. Sensors 2017, 17, 850. [Google Scholar] [CrossRef] [PubMed]
  38. Abdullah, K.A.; Batal, M.A.; Hamdan, R.; Khalil, T.; Salame, C. Noise effect on performance of IR PVDF pyroelectric detector. AIP Conf. Proc. 2018, 1968, 030079. [Google Scholar] [CrossRef]
  39. Goel, T.C.; Kumar, P.; James, A.R.; Prakash, C. Processing and Dielectric Properties of Sol-Gel Derived PMN-PT (68:32) Thin Films. J. Electroceram. 2004, 13, 503–507. [Google Scholar] [CrossRef]
  40. Wang, J.; Wong, K.H.; Chan, H.L.W.; Choy, C.L. Composition control and electrical properties of PMN-PT thin films around the morphotropic boundary. Appl. Phys. A 2004, 79, 551–556. [Google Scholar] [CrossRef]
  41. Savage, A. Pyroelectricity and Spontaneous Polarization in LiNbO3. J. Appl. Phys. 1966, 37, 3071–3072. [Google Scholar] [CrossRef]
  42. Brick, D.; Emre, E.; Grossmann, M.; Dekorsy, T.; Hettich, M. Picosecond Photoacoustic Metrology of SiO2 and LiNbO3 Layer Systems Used for High Frequency Surface-Acoustic-Wave Filters. Appl. Sci. 2017, 7, 822. [Google Scholar] [CrossRef]
  43. Chang, H.H.S.; Huang, Z. Laminate composites with enhanced pyroelectric effects for energy harvesting. Smart Mater. Struct. 2010, 19, 065018. [Google Scholar] [CrossRef]
  44. Lin, D.; Li, Z.; Li, F.; Zhang, S. Direct observation of domain wall motion and novel dielectric loss in 0.23Pb(In1/2Nb1/2)O3–0.42Pb(Mg1/3Nb2/3)O3–0.35PbTiO3 crystals. CrystEngComm 2013, 15, 6292–6296. [Google Scholar] [CrossRef]
  45. Dabin, L.; Zhenrong, L.; Zhuo, X.; Xi, Y. The polarization fatigue behavior in Pb(Mg1/3Nb2/3)O3-0.32PbTiO3 single crystals. J. Phys. Conf. Ser. 2009, 152, 012088. [Google Scholar]
Figure 1. XRD patterns of 2.5% Pr3+ doped Pb(Mg1/3Nb2/3)O3-0.30PbTiO3 (Pr-PMN-PT) thin films as a function of annealing temperatures between 600 °C and 700 °C: (a,b); XRD patterns of 2.5% Pr-PMN-PT thin films annealed at 650 °C as a function of annealing times between 1 min and 5 min: (c,d).
Figure 1. XRD patterns of 2.5% Pr3+ doped Pb(Mg1/3Nb2/3)O3-0.30PbTiO3 (Pr-PMN-PT) thin films as a function of annealing temperatures between 600 °C and 700 °C: (a,b); XRD patterns of 2.5% Pr-PMN-PT thin films annealed at 650 °C as a function of annealing times between 1 min and 5 min: (c,d).
Materials 11 02392 g001
Figure 2. (a) FE-SEM images of surface morphologies for 2.5% Pr-PMN-PT thin films annealed at 650 °C for 2 min, (b) FE-SEM images of cross-section of 2.5% Pr-PMN-PT thin films, (c) grain size distributions of 2.5% Pr-PMN-PT thin films, (d) 2D piezoresponse force microscope (PFM) images of surface morphologies for 2.5% Pr-PMN-PT thin films, (e) 3D atomic force microscope (AFM) images of surface morphologies for 2.5% Pr-PMN-PT thin films, and (f) a schematic illustration of the device for electric properties measurement.
Figure 2. (a) FE-SEM images of surface morphologies for 2.5% Pr-PMN-PT thin films annealed at 650 °C for 2 min, (b) FE-SEM images of cross-section of 2.5% Pr-PMN-PT thin films, (c) grain size distributions of 2.5% Pr-PMN-PT thin films, (d) 2D piezoresponse force microscope (PFM) images of surface morphologies for 2.5% Pr-PMN-PT thin films, (e) 3D atomic force microscope (AFM) images of surface morphologies for 2.5% Pr-PMN-PT thin films, and (f) a schematic illustration of the device for electric properties measurement.
Materials 11 02392 g002
Figure 3. Element mapping of 2.5% Pr-PMN-PT thin films annealed at 650 °C for 2 min.
Figure 3. Element mapping of 2.5% Pr-PMN-PT thin films annealed at 650 °C for 2 min.
Materials 11 02392 g003
Figure 4. (a) P-E hysteresis loops for pure PMN-PT and Pr-PMN-PT thin films annealed at 650 °C for 2 min, (b) leakage current density dependent on external electric field for pure PMN-PT and 2.5% Pr-PMN-PT thin films.
Figure 4. (a) P-E hysteresis loops for pure PMN-PT and Pr-PMN-PT thin films annealed at 650 °C for 2 min, (b) leakage current density dependent on external electric field for pure PMN-PT and 2.5% Pr-PMN-PT thin films.
Materials 11 02392 g004
Figure 5. (a) Dielectric coefficient (εr) and dielectric loss (tanδ) as a function of frequency for 2.5% Pr-PMN-PT thin films annealed at 650 °C for 2 min, (b) dielectric coefficient and dielectric loss as a function of temperature for 2.5% Pr-PMN-PT thin films at 100 Hz and 1 kHz.
Figure 5. (a) Dielectric coefficient (εr) and dielectric loss (tanδ) as a function of frequency for 2.5% Pr-PMN-PT thin films annealed at 650 °C for 2 min, (b) dielectric coefficient and dielectric loss as a function of temperature for 2.5% Pr-PMN-PT thin films at 100 Hz and 1 kHz.
Materials 11 02392 g005
Figure 6. (a) Pyroelectric coefficient (py) dependent of Pr content in Pr-PMN-PT thin film annealed at 650 °C for 2 min; (b) pyroelectric coefficient (py) measured in pure PMN-PT and 2.5% Pr-PMN-PT thin films at temperature between −50 °C and 100 °C.
Figure 6. (a) Pyroelectric coefficient (py) dependent of Pr content in Pr-PMN-PT thin film annealed at 650 °C for 2 min; (b) pyroelectric coefficient (py) measured in pure PMN-PT and 2.5% Pr-PMN-PT thin films at temperature between −50 °C and 100 °C.
Materials 11 02392 g006
Table 1. Dielectric, pyroelectric, and ferroelectric properties in (001)-oriented PMN-PT based ferroelectric thin films at room temperature.
Table 1. Dielectric, pyroelectric, and ferroelectric properties in (001)-oriented PMN-PT based ferroelectric thin films at room temperature.
Thin FilmsSubstratesThickness
(nm)
Methodεr
@1kHz
tanδ
(%)
py
(μC/m2K)
pr
(μC/cm2)
Reference
PMN-0.30PTPt/TiO2/SiO2/Si450Sol-gel11403.17511.5This work
2.5%Pr-PMN-0.30PTPt/TiO2/SiO2/Si450Sol-gel24001.516717.3This work
PMN-0.30PTPt/TiO2/SiO2/Si350CSD17100.8//[20]
PMN-0.32PTPt/Si/ITO glass800Sol-gel1110//0.55[39]
PMN-0.30PTSrRuO3/SrTiO3200CSD~2600~5/~13[19]
PMN-0.32PTBa0.5Sr0.5RuO3/
NdScO3
150PLD~1200~7−550~8[4]
PMN-0.32PTLa0.5Sr0.5CoO3/
LaAlO3
150PLD~2500~7160–300~11[40]
LiNbO3SiO2325 × 103///40~50[41,42]
PVDFPMMA28 × 103/~5~223/[37,43]
εr dielectric constant, tanδ dielectric loss, py pyroelectric coefficient, pr remnant polarization, pulsed-laser deposition (PLD), chemical solution deposition (CSD).

Share and Cite

MDPI and ACS Style

Cai, C.; Zhang, D.; Liu, W.; Wang, J.; Zhou, S.; Su, Y.; Sun, X.; Lin, D. Synthesis, Giant Dielectric, and Pyroelectric Response of [001]-Oriented Pr3+ Doped Pb(Mg1/3Nb2/3)O3-PbTiO3 Ferroelectric Nano-Films Grown on Si Substrates. Materials 2018, 11, 2392. https://doi.org/10.3390/ma11122392

AMA Style

Cai C, Zhang D, Liu W, Wang J, Zhou S, Su Y, Sun X, Lin D. Synthesis, Giant Dielectric, and Pyroelectric Response of [001]-Oriented Pr3+ Doped Pb(Mg1/3Nb2/3)O3-PbTiO3 Ferroelectric Nano-Films Grown on Si Substrates. Materials. 2018; 11(12):2392. https://doi.org/10.3390/ma11122392

Chicago/Turabian Style

Cai, Changlong, Deqiang Zhang, Weiguo Liu, Jun Wang, Shun Zhou, Yongming Su, Xueping Sun, and Dabin Lin. 2018. "Synthesis, Giant Dielectric, and Pyroelectric Response of [001]-Oriented Pr3+ Doped Pb(Mg1/3Nb2/3)O3-PbTiO3 Ferroelectric Nano-Films Grown on Si Substrates" Materials 11, no. 12: 2392. https://doi.org/10.3390/ma11122392

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