Next Article in Journal
Measuring Roadway Lane Widths Using Connected Vehicle Sensor Data
Previous Article in Journal
Accuracy of Inertial Measurement Units When Applied to the Countermovement Jump of Track and Field Athletes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Smart Sensing Multifunctionalities Based on Barium Strontium Titanate Thin Films

1
College of Physics and Information Engineering, Fuzhou University, Fuzhou 350108, China
2
FZU-Jinjiang Joint Institute of Microelectronics, Jinjiang Science and Education Park, Fuzhou University, Jinjiang 362200, China
3
School of Materials Science and Engineering, Nanyang Technological University, 50 Nanyang Avenue, Singapore 639798, Singapore
4
Academy of Hi-Tech Research, Hunan Institute of Traffic Engineering, Hengyang 421099, China
*
Authors to whom correspondence should be addressed.
Sensors 2022, 22(19), 7183; https://doi.org/10.3390/s22197183
Submission received: 30 August 2022 / Revised: 17 September 2022 / Accepted: 20 September 2022 / Published: 22 September 2022
(This article belongs to the Section Sensor Materials)

Abstract

:
Sensors that have low power consumption, high scalability and the ability of rapidly detecting multitudinous external stimulus are of great value in cyber-physical interactive applications. Herein, we reported the fabrication of ferroelectric barium strontium titanate ((Ba70Sr30)TiO3, BST) thin films on silicon substrates by magnetron sputtering. The as-grown BST films have a pure perovskite structure and exhibit excellent ferroelectric characteristics, such as a remnant polarization of 2.4 μC/cm2, a ferro-to-paraelectric (tetragonal-to-cubic) phase transition temperature of 31.2 °C, and a broad optical bandgap of 3.58 eV. Capacitor-based sensors made from the BST films have shown an outstanding average sensitivity of 0.10 mV·Pa−1 in the 10–80 kPa regime and work extremely steadily over 1000 cycles. More importantly, utilizing the Pockels effect, optical manipulation in BST can be also realized by a smaller bias and its electro-optic coefficient reff is estimated to be 83.5 pmV−1, which is 2.6 times larger than in the current standard material (LiNbO3) for electro-optical devices. Our work established BST thin film as a powerful design paradigm toward on-chip integrations with diverse electronics into sensors via CMOS-comparable technique.

1. Introduction

With the rapid development of the Internet of Things (IoT), a gigantic array of electronics and sensors have been quietly functioning in our daily life, allowing for signal processing, wireless communication, edge computing and many more [1,2,3]. In particular, the sensors are the core components that directly enable interconversion between the external stimuli and the signals. In recent years, great efforts have been dedicated to achieving highly selective and highly scalable sensors. Among them, both capacitive- and transistor-based sensors have attracted considerable interest, owing to their simple designs, lower power consumption, and complementary metal oxide semiconductor (CMOS)-comparable manufacture technique [4,5,6,7]. They can be integrated with a diverse range of electronics into sensors for monitoring optical, mechanical, thermal and vital parameters, thereby demonstrating the great potential of numerous existing and emerging applications, such as human–machine interfaces, robotic technologies, and wearable healthcare monitors.
Noticeably, lead-free ferroelectric materials have emerged as a viable platform for multimodal distributed sensing functionalities, owing to their high capacitance, switchable spontaneous polarization, pyroelectric, piezoelectric, photovoltaic, and electro-optic effects [3,8,9,10,11]. The unique characteristics of ferroelectrics can be utilized to measure pressure, force, temperature, humidity, and IR radiation [12,13]. Tian et al. investigated reversible bulk polarization switching induced by construction of chemical bonds at the surface of ferroelectric BiFeO3 in aqueous solution, opening up new opportunities for chemical stimuli sensing [14]. Additionally, Yang group designed a vertical ITO/BaTiO3/ITO photodetector, which generates a dramatic increase of photocurrent, revealing the potential for optical monitoring [15]. However, aforementioned works either adopt bulk ferroelectric materials or single crystal substrate, such as SrTiO3, which cannot simultaneously achieve both lower power consumption and CMOS-comparable technology. Currently, the fabrication of ferroelectric thin films directly grown on Si wafer remains challenging, because most perovskite materials, such as (Pb,Zr)TiO3, BiFeO3 and Pb(Mg2/3Nb1/3)O3-PbTiO3, require a certain degree of lattice matching with the substrate to obtain a pure perovskite phase [14,16,17]. In this letter, we reported the growth of high-quality ferroelectric BST thin films on Si substrates by magnetron sputtering. Subsequently, the formation mechanism of perovskite phase, spontaneous polarization and temperature-dependent dielectric characteristics were investigated. Additionally, capacitor-based BST sensors have exhibited a remarkable sensitivity of 0.10 mV·Pa−1 in the range from 10 kPa to 80 kPa, and even work extremely steadily on over 1000 cycles. Optical performances of BST thin films can be easily manipulated by a smaller external voltage, indicating the great potential for the current standard material (LiNbO3) replacement. Owing to their excellent ferroelectric properties, mechanical and optical manipulation in the resulting BST films were realized, providing a new opportunity for multifunctional sensing applications.

2. Experimental Section

BST thin films were synthesized at a substrate temperature of around 300 °C and at a mixed Ar/O2 gas atmosphere by magnetron sputtering, followed by a post-annealing process at 680 °C for 10 min. During the deposition, the pressure is about 1 Torr and the power is 150 W. The resulting BST thin film is calculated to be 500 nm. X-ray diffraction (XRD, 20 kV) with Cu kα anode and scanning electron microscopy (SEM, 5 kV) equipped with energy-dispersive X-ray spectroscopy was carried out to investigate the crystallinity and the component of the as-grown BST thin films. The polarization–electric field (P-E) loop and temperature-dependent dielectric characteristics were measured by a high-precision LCR meter and Radiant precision multiferroic II. As for the mechanical manipulation, one setup was used to record the output voltage generated by BST films under the various pressures. Transmittance spectrum and the refractive index of all samples were characterized using a UV–Vis–NIR spectrometer and an ellipsometry. Lastly, a customer-designed facility equipped with a laser, two polarizers, an analyzer, a photodetector and a power amplifier were used to realize the optical manipulation.

3. Results and Discussion

Figure 1a shows a typical XRD pattern of perovskite-structural ferroelectric thin film, without the presence of any intermediate phases such as pyrochlore phase [17]. Since BST materials have a tetragonal P4mm symmetry, its lattice parameter a and c can be derived from XRD results (a = 3.9587 Å and c = 4.0313 Å), which is slightly larger than that of BST epitaxially grown on single crystal MgO substrate by molecular chemical vapor deposition (a = 3.9100 Å, c = 4.0030 Å) and very close to that of poly-crystalline BST on Pt by sputtering (a = 3.9870 Å, c = 4.0030 Å) [18,19]. It is mainly attributed to the thermal expansion mismatch of the substrates with Si. It is reported that the thermal expansion coefficient of Si is approximately 2.6 × 10−6 °C−1, which is far less than 6 × 10−6 °C−1 for BST and 12.6 × 10−6 °C−1 for MgO [19,20,21]. The other factor could be the lattice mismatch between the film and the Si substrate (a = 5.4300 Å). Energy dispersive X-ray spectroscopy (EDX) analysis, as shown in Figure 1b, detected four elements: Ba, Sr, Ti, and O, indicating the presence of BST thin films. More importantly, the percentage of Ti and O is 61.02:21.04 while that of Ba and Sr is 12.41:5.53 (The inset table), which is consistent with 3:1 for Ti/O and 7:3 for Ba/Sr in raw materials. The accurate and stable stoichiometric proportion in these samples is desirable for ferroelectric applications.
The X-ray photoelectron spectroscopy (XPS) technique is further performed to identify the elements and chemical states of the as-synthesized films. Figure 2a shows a XPS survey of BST thin film, which records C, Ba, Sr, Ti, and O. Subsequently, the high-resolution XPS spectra of C 1s, Ba 3d, Sr 3d, Ti 2p, and O 1s peaks in BST thin film are shown in Figure 2b–f. The C 1s at 284.6 eV can be used as a reference for the systematic shifts of the spectra (Figure 2b) [22]. Gaussian fitting analysis reveals that Ba 3d can be fitted by two peaks at 779.5 eV for Ba 3d3/2 and 794.9 eV for Ba 3d5/2, which are assigned to Ba atoms in the BST perovskite phase [23]. Similarly, two fitting peaks can be found in the spectra of both Sr 3d and Ti 2p. The binding energies of the former are 132.8 eV and 134.6 eV, while 458.0 eV and 463.6 eV in the latter. These results further confirm that Sr and Ti atoms relate to Sr2+ and Ti4+ in perovskites [24,25]. Furthermore, O 1s produces two peaks as shown in Figure 2f. The binding energy of 529.8 eV is corresponding to the binding energy of O atoms in the perovskite phase. Another peak at 531.6 eV is ascribed to the non-lattice oxygen due to the surface absorption [26]. Thus, XPS analysis confirms the formation of the perovskite phase in the view of element chemical bonding, which is consistent with the results of XRD and SEM results.
The ferroelectricity of BST thin films is demonstrated by the polarization-electric field (P-E) hysteresis loop. Figure 3a exhibits an asymmetric and slim P-E curve with the remnant polarization (Pr) of 2.4 µC/cm2, compared to the values of 2.5 µC/cm2 for BST by pulsed laser deposition and 2.98 µC/cm2 for Zr-doped BST by sol-gel method [27,28]. Figure 3b shows the dielectric constant (εr) and the loss (tan δ) versus temperature trend at a diverse frequency (0.1 kHz−1000 kHz). Notably, the typical ferroelectric materials follow the Curie-Weiss law above Curie temperature, which is expressed by: 1/ε = (TTc)/C, where Tc is the Curie temperature and C is the constant. Therefore, a ferro-to-paraelectric (tetragonal-to-cubic) phase transition occurs at 31.2 °C, corresponded to the Curie temperature of BST. As the temperature ramps up, the εr initially decreases and is followed by rapid increasing. Meanwhile, the corresponding dielectric loss remains largely constant below 400 °C and then increases sharply at high temperatures. Similar phenomenon was also observed in BST bulk ceramics and (Pb,La)(Zr,Ti)O3 relaxor [29,30].
Mechanical manipulation in BST thin film is realized through a customer-designed testing system. The details on experimental sections can be found in our previous work [3]. Figure 4a shows the piezoelectric voltage response induced by applying a periodic mechanical pressure of 10 kPa. Notably, the output voltage demonstrates the similar periodicity, with a reliable maximum value of ~0.9 V. This can be mainly attributed to the piezoelectric effect of BST. As illustrated in Figure 4b, during the compression of the BST which is sandwiched between two electrodes, electrons are injected from the top electrode to the BST surface. During the subsequent releasing stage, charges transfer to the bottom electrode. Thus, the structure established the voltage potential between two electrodes, allowing electrons to flow through the top electrode to the bottom electrode, maintaining the electrostatic equilibrium. Generally, the relationship between the output voltage (V) and the applied pressure (P) is given by [31]:
V = (hd33P)/(ε0εr)
where h, d33, ε0, εr refers to the film thickness, the piezoelectric constant of BST, the permittivity of free space, and the relative dielectric constant of BST, respectively. As the mechanical pressure varies from 10 kPa to 80 kPa, a varying output voltage is generated, which is plotted in Figure 4c. In this case, the measured output voltages of BST thin film increase linearly in a range of 0.9 V to 8.6 V, which is in agreement with Equation (1) (Figure 4d). The sensitivity (δ) of the BST-based pressure sensors is defined as: δ = V/P, where V and P are the output voltage and the applied pressure. In the regime from 10 kPa to 80 kPa, the average sensitivity δ is calculated to be 0.10 mV·Pa−1, which is considerably higher than the 0.05–0.07 mV·Pa−1 range achieved by other ferroelectric sensors [32,33]. More importantly, almost no degradation in performance was observed, even after 1000 cycles (the inset of Figure 4d), indicating excellent operation stability of our BST-based sensing device.
Figure 5a displays the typical transmittance curve of the BST film measured with a UV-Vis Spectrometer, indicating a high transparency under visible to near-infrared light. The inset is a plot of the absorption edge, which can be used to derive the optical band gap of BST using: (αhυ)2 = C(Eg), where α, , and Eg are the absorption coefficient, the photon energy, and the optical band gap, respectively. From the inset, the optical band gap of BST is estimated to be 3.58 eV, which is consistent with the data reported in the literature [34,35]. The measured ellipsometric parameters Tan Ψ and Cos Δ were applied, respectively, to derive the refractive index (n) of BST films (Inset of Figure 5b). In this case, the n versus the wavelength (λ) can be expressed as: n(λ) = A + (B/λ2) + (C/λ4), where A, B, and C are the Cauchy parameters [16]. The as-grown BST sample exhibits a refractive index of 2.42 at the wavelength of 632 nm (Figure 5b). This value is comparable to 2.43 for Pb(Zr,Ti)O3 and 2.34 for (Pb,La)(Zr,Ti)O3 and slightly smaller than that of Pb(Mg2/3Nb1/3)O3-PbTiO3 (PMN-PT, n = 2.599) [36,37,38]. The refractive index in perovskites can be easily manipulated by external stimulus such as the electric field, e.g., birefringence, which in turn provides the opportunity to manipulate the optical performances of BST [39].
In addition to mechanical sensitivity, the optical manipulation in BST films can be also realized. Figure 6a demonstrates the laser intensity modulation under diverse applied voltages. Notably, the relationship between the intensity and the applied voltage is linear, evidenced by Figure 6b. Utilizing a modified Senarmont compensator style, the inset of Figure 6b is the schematics of experimental setup, equipped with a stabilized He–Ne laser, a polarizer, an analyzer, and a photodetector [40]. The output intensity (I), recorded by the photodetector, can be given by: I = 1/2 I0(1 + sinΓ) = I0(1 + sin(2πlΔn/λ)), where I0 is the initial intensity of the laser, Γ is the phase retardation, l is the thickness of BST, and Δn is the birefringence variation under the applied electric field. As we know, the electric field E is applied perpendicular to the ferroelectric film, leading to a shift in the refractive index. Thus, the field dependence E of the index n(E) is given by [39]:
n(E) = n − 1/2 rn3E + 1/2 ξn2E2
where r and ξ are the Pockels and Kerr effect coefficients, respectively. Tetragonal BST is a positive uniaxial optical crystal with its optical axis along the polarization direction. When applying an E1 or E2 electric field along the (110) direction, for example, using the coordinate system shown in Figure 6c, the spontaneous polarization inside the (110) plane was poled along the Od1 and Od2 directions. Figure 6d displays the refractive index ellipse in BST, showing the refractive index nx and nz along different crystalline directions (left). Applying an electric field E changes the birefringence (Δn) and rotates the optical axis depending on the orientation of E. As for birefringent materials such as BST, nx = ny = no and nz = ne, where no and ne are the ordinary and extraordinary refractive indices of the medium, respectively [41]. Therefore, the modulation simplifies in Equation (2) to no(E) = no − 1/2 r13no3E or ne(E) = ne − 1/2 r13ne3E. When defining an effective electro-optic coefficient as reff = r33r13(no3/ne3), the birefringence (Δn) can be given by: Δn = neno = 1/2 reffn3E. Generally, the change in refractive index induced by the electric field is very small for direct measurements. Therefore, the change in refractive index is always replaced by an optical path difference or phase change in practical experiments [39]. Herein, the electro-optic coefficient reff of BST films is estimated to be 83.5 pmV−1. It is noteworthy that this value is also comparable with the commonly used transverse BST coefficient in (111)-oriented BST (87.8 pmV−1) and is far superior to that of LiNbO3 crystals (31.8 pmV−1), opening up new avenues to replace LiNbO3 for enabling future high-speed and cost-effective optical communication networks [42,43].

4. Conclusions

To sum, barium strontium titanate thin films were directly deposited on silicon substrates by sputtering technique. Their crystal phase, spontaneous polarization and temperature-dependent dielectric characteristics were subsequently studied. In addition to the high transmittance over a wide wavelength regime, the as-grown BST films have exhibited a remnant polarization of 2.4 μC/cm2, a ferro-to-paraelectric (tetragonal-to-cubic) phase transition temperature of 31.2 °C, and a broad optical bandgap of 3.58 eV. Owing to their excellent ferroelectric performances, the resulted BST films exhibit a highly sensitive pressure sensor with an average sensitivity of 0.10 mV·Pa−1 in the 10–80 kPa regime and excellent operation stability of 1000 cycles or above. Additionally, they can be also used to manipulate optical properties simply by applying a smaller voltage, which is linked to the electric field-induced Pockels effect. An electro-optic coefficient of 83.5 pmV−1 is observed. This study paves the way for a new opportunity for on-chip integrations with various electronics into sensors via CMOS-comparable technique, allowing for low power consumption and multimodel functionalities.

Author Contributions

Conceptualization, M.Z. and L.W.; methodology, M.Z.; Data curation, C.W., Y.S. and L.G.; formal analysis, M.Z., L.W. and Y.B.; writing—review and editing, M.Z., L.W. and Y.Z.; supervision, M.Z. and Y.B. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledge the support from Fujian Provincial Science and Technology Planning Project (No. 2022I0006).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors have no conflict of interest to disclose.

References

  1. Boutry, C.M.; Kaizawa, Y.; Schroeder, B.C.; Chortos, A.; Legrand, A.; Wang, Z.; Chang, J.; Fox, P.; Bao, Z. A Stretchable and Biodegradable Strain and Pressure Sensor for Orthopaedic Application. Nat. Electron. 2018, 1, 314–321. [Google Scholar] [CrossRef]
  2. Stadlober, B.; Zirkl, M.; Irimia-Vladu, M. Route towards Sustainable Smart Sensors: Ferroelectric Polyvinylidene Fluoride-Based Materials and Their Integration in Flexible Electronics. Chem. Soc. Rev. 2019, 48, 1787–1825. [Google Scholar] [CrossRef] [PubMed]
  3. Zhu, M.; Chng, S.S.; Cai, W.; Liu, C.; Du, Z. Piezoelectric Polymer Nanofibers for Pressure Sensors and Their Applications in Human Activity Monitoring. RSC Adv. 2020, 10, 21887–21894. [Google Scholar] [CrossRef] [PubMed]
  4. Li, X.; Fan, Y.J.; Li, H.Y.; Cao, J.W.; Xiao, Y.C.; Wang, Y.; Liang, F.; Wang, H.L.; Jiang, Y.; Wang, Z.L.; et al. Ultracomfortable Hierarchical Nanonetwork for Highly Sensitive Pressure Sensor. ACS Nano 2020, 14, 9605–9612. [Google Scholar] [CrossRef] [PubMed]
  5. Huang, Y.-C.; Liu, Y.; Ma, C.; Cheng, H.-C.; He, Q.; Wu, H.; Wang, C.; Lin, C.-Y.; Huang, Y.; Duan, X. Sensitive Pressure Sensors Based on Conductive Microstructured Air-Gap Gates and Two-Dimensional Semiconductor Transistors. Nat. Electron. 2020, 3, 59–69. [Google Scholar] [CrossRef]
  6. Persano, L.; Dagdeviren, C.; Su, Y.; Zhang, Y.; Girardo, S.; Pisignano, D.; Huang, Y.; Rogers, J.A. High Performance Piezoelectric Devices Based on Aligned Arrays of Nanofibers of Poly(Vinylidenefluoride-Co-Trifluoroethylene). Nat. Commun. 2013, 4, 1633. [Google Scholar] [CrossRef] [PubMed]
  7. Khan, A.I.; Chatterjee, K.; Wang, B.; Drapcho, S.; You, L.; Serrao, C.; Bakaul, S.R.; Ramesh, R.; Salahuddin, S. Negative Capacitance in a Ferroelectric Capacitor. Nat. Mater. 2015, 14, 182–186. [Google Scholar] [CrossRef]
  8. Yao, K.; Chen, S.; Lai, S.C.; Yousry, Y.M. Enabling Distributed Intelligence with Ferroelectric Multifunctionalities. Adv. Sci. 2022, 9, 2103842. [Google Scholar] [CrossRef]
  9. You, L.; Zheng, F.; Fang, L.; Zhou, Y.; Tan, L.Z.; Zhang, Z.; Ma, G.; Schmidt, D.; Rusydi, A.; Wang, L.; et al. Enhancing Ferroelectric Photovoltaic Effect by Polar Order Engineering. Sci. Adv. 2018, 4, 1–10. [Google Scholar] [CrossRef]
  10. Mazon, T.; Zaghete, M.A.; Varela, J.A.; Longo, E. Barium Strontium Titanate Nanocrystalline Thin Films Prepared by Soft Chemical Method. J. Eur. Ceram. Soc. 2007, 27, 3799–3802. [Google Scholar] [CrossRef]
  11. Han, W.; Kim, T.; Yoo, B.; Park, H.-H. Tunable Dielectric Properties of Poly(Vinylidenefluoride-Co-Hexafluoropropylene) Films with Embedded Fluorinated Barium Strontium Titanate Nanoparticles. Sci. Rep. 2018, 8, 4086. [Google Scholar] [CrossRef] [PubMed]
  12. Damjanovic, D.; Muralt, P.; Setter, N. Ferroelectric Sensors. IEEE Sens. J. 2001, 1, 191–206. [Google Scholar] [CrossRef]
  13. Chen, S.; Yuan, S.; Hou, Z.; Tang, Y.; Zhang, J.; Wang, T.; Li, K.; Zhao, W.; Liu, X.; Chen, L.; et al. Recent Progress on Topological Structures in Ferroic Thin Films and Heterostructures. Adv. Mater. 2021, 33, 2000857. [Google Scholar] [CrossRef] [PubMed]
  14. Tian, Y.; Wei, L.; Zhang, Q.; Huang, H.; Zhang, Y.; Zhou, H.; Ma, F.; Gu, L.; Meng, S.; Chen, L.-Q.; et al. Water Printing of Ferroelectric Polarization. Nat. Commun. 2018, 9, 3809. [Google Scholar] [CrossRef]
  15. Ma, N.; Yang, Y. Boosted Photocurrent in Ferroelectric BaTiO3 Materials via Two Dimensional Planar-Structured Contact Configurations. Nano Energy 2018, 50, 417–424. [Google Scholar] [CrossRef]
  16. Zhu, M.; Du, Z.; Jing, L.; Yoong Tok, A.I.; Tong Teo, E.H. Optical and Electro-Optic Anisotropy of Epitaxial PZT Thin Films. Appl. Phys. Lett. 2015, 107, 031907. [Google Scholar] [CrossRef]
  17. Du, Z.H.; Zhang, T.S.; Zhu, M.M.; Ma, J. Perovskite Crystallization Kinetics and Dielectric Properties of the PMN-PT Films Prepared by Polymer-Modified Sol-Gel Processing. J. Mater. Res. 2009, 24, 1576–1584. [Google Scholar] [CrossRef]
  18. Funakubo, H.; Takeshima, Y.; Nagano, D.; Shinozaki, K.; Mizutani, N. Crystal Structure and Dielectric Property of Epitaxially Grown (Ba,Sr)TiO3 Thin Film Prepared by Molecular Chemical Vapor Deposition. J. Mater. Res. 1998, 13, 3512–3518. [Google Scholar] [CrossRef]
  19. Park, W.Y.; Park, M.H.; Lee, J.H.; Yoon, J.H.; Han, J.H.; Choi, J.-H.; Hwang, C.S. Strain Evolution of Each Type of Grains in Poly-Crystalline (Ba,Sr)TiO3 Thin Films Grown by Sputtering. Sci. Rep. 2012, 2, 939. [Google Scholar] [CrossRef] [PubMed]
  20. Engberg, C.J.; Zehms, E.H. Thermal Expansion of Al2O3, BeO, MgO, B4C, SiC, and TiC Above 1000 °C. J. Am. Ceram. Soc. 1959, 42, 300–305. [Google Scholar] [CrossRef]
  21. He, Y. Heat Capacity, Thermal Conductivity, and Thermal Expansion of Barium Titanate-Based Ceramics. Thermochim. Acta 2004, 419, 135–141. [Google Scholar] [CrossRef]
  22. Zhang, Z.; Zhao, Y.; Zhu, M. NiO Films Consisting of Vertically Aligned Cone-Shaped NiO Rods. Appl. Phys. Lett. 2006, 88, 033101. [Google Scholar] [CrossRef]
  23. Craciun, V.; Singh, R.K. Characteristics of the Surface Layer of Barium Strontium Titanate Thin Films Deposited by Laser Ablation. Appl. Phys. Lett. 2000, 76, 1932–1934. [Google Scholar] [CrossRef]
  24. Yan, H.; Wang, J.; Yang, B.; Wang, S.; Xing, H.; Jin, K. Tunability of Band-Gap in Barium Strontium Titanate Films on Anodic Aluminum Oxide Templates. Thin Solid Films 2018, 649, 187–191. [Google Scholar] [CrossRef]
  25. Zhu, M.; Zhang, H.; Favier, S.W.L.; Zhao, Y.; Guo, H.; Du, Z. A General Strategy towards Controllable Replication of Butterfly Wings for Robust Light Photocatalysis. J. Mater. Sci. Technol. 2022, 105, 286–292. [Google Scholar] [CrossRef]
  26. Zhu, M.; Zhang, H.; Du, Z.; Liu, C. Structural Insight into the Optical and Electro-Optic Properties of Lead Zirconate Titanate for High-Performance Photonic Devices. Ceram. Int. 2019, 45, 22324–22330. [Google Scholar] [CrossRef]
  27. Ortega, N.; Kumar, A.; Resto, O.; Maslova, O.A.; Yuzyuk, Y.I.; Scott, J.F.; Katiyar, R.S. Compositional Engineering of BaTiO3/(Ba,Sr)TiO3 Ferroelectric Superlattices. J. Appl. Phys. 2013, 114, 104102. [Google Scholar] [CrossRef]
  28. Mahmood, A.; Naeem, A.; Iqbal, Y.; Ullah, A. Dielectric and Ferroelectric Properties of the Sol-Gel-Derived Zr-Doped Ba0.7Sr0.3TiO3 Polycrystalline Ceramic Systems. Int. J. Appl. Ceram. Technol. 2017, 14, 604–610. [Google Scholar] [CrossRef]
  29. Ianculescu, A.; Berger, D.; Viviani, M.; Ciomaga, C.E.; Mitoseriu, L.; Vasile, E.; Drăgan, N.; Crişan, D. Investigation of Ba1−xSrxTiO3 Ceramics Prepared from Powders Synthesized by the Modified Pechini Route. J. Eur. Ceram. Soc. 2007, 27, 3655–3658. [Google Scholar] [CrossRef]
  30. Kang, B.S.; Choi, S.K.; Park, C.H. Diffuse Dielectric Anomaly in Perovskite-Type Ferroelectric Oxides in the Temperature Range of 400–700 °C. J. Appl. Phys. 2003, 94, 1904–1911. [Google Scholar] [CrossRef] [Green Version]
  31. Surmenev, R.A.; Orlova, T.; Chernozem, R.V.; Ivanova, A.A.; Bartasyte, A.; Mathur, S.; Surmeneva, M.A. Hybrid Lead-Free Polymer-Based Nanocomposites with Improved Piezoelectric Response for Biomedical Energy-Harvesting Applications: A Review. Nano Energy 2019, 62, 475–506. [Google Scholar] [CrossRef]
  32. Ye, S.; Cheng, C.; Chen, X.; Chen, X.; Shao, J.; Zhang, J.; Hu, H.; Tian, H.; Li, X.; Ma, L.; et al. High-Performance Piezoelectric Nanogenerator Based on Microstructured P(VDF-TrFE)/BNNTs Composite for Energy Harvesting and Radiation Protection in Space. Nano Energy 2019, 60, 701–714. [Google Scholar] [CrossRef]
  33. Sekine, T.; Sugano, R.; Tashiro, T.; Sato, J.; Takeda, Y.; Matsui, H.; Kumaki, D.; Dos Santos, F.D.; Miyabo, A.; Tokito, S. Fully Printed Wearable Vital Sensor for Human Pulse Rate Monitoring Using Ferroelectric Polymer. Sci. Rep. 2018, 8, 1–10. [Google Scholar]
  34. Sharma, P.K.; Messing, G.L.; Agrawal, D.K. Structural, Ultraviolet-Shielding, Dielectric and Ferroelectric Properties of Ba1−xSrxTiO3 (X = 0.35) Thin Films Prepared by Sol-Gel Method in Presence of Pyrrole. Thin Solid Films 2005, 491, 204–211. [Google Scholar] [CrossRef]
  35. Boubaia, A.; Assali, A.; Berrah, S.; Bennacer, H.; Zerifi, I.; Boukortt, A. Band Gap and Emission Wavelength Tuning of Sr-Doped BaTiO3 (BST) Perovskites for High-Efficiency Visible-Light Emitters and Solar Cells. Mater. Sci. Semicond. Process. 2021, 130, 105837. [Google Scholar] [CrossRef]
  36. Zhu, M.; Du, Z.; Li, H.; Chen, B.; Jing, L.; Tay, R.Y.J.; Lin, J.; Tsang, S.H.; Teo, E.H.T. Tuning Electro-Optic Susceptibity via Strain Engineering in Artificial PZT Multilayer Films for High-Performance Broadband Modulator. Appl. Surf. Sci. 2017, 425, 1059–1065. [Google Scholar] [CrossRef]
  37. Du, Z.H.; Zhang, T.S.; Zhu, M.M.; Ma, J. Growth Mode of Sol-Gel Derived PLT Seeding Layers on Glass Substrates and Its Effect on Templating the Oriented Growth of PLZT Thin Films. J. Appl. Phys. 2009, 105, 061612. [Google Scholar] [CrossRef]
  38. Wan, X.; Luo, H.; Zhao, X.; Wang, D.Y.; Chan, H.L.W.; Choy, C.L. Refractive Indices and Linear Electro-Optic Properties of (1-x)Pb(Mg1/3Nb2/3)O3-XPbTiO3 Single Crystals. Appl. Phys. Lett. 2004, 85, 5233–5235. [Google Scholar] [CrossRef]
  39. Wessels, B.W. Ferroelectric Epitaxial Thin Films for Integrated Optics. Annu. Rev. Mater. Res. 2007, 37, 659–679. [Google Scholar] [CrossRef]
  40. Dupuy, A.D.; Kodera, Y.; Garay, J.E. Unprecedented Electro-Optic Performance in Lead-Free Transparent Ceramics. Adv. Mater. 2016, 28, 7970–7977. [Google Scholar] [CrossRef]
  41. Zhu, M.; Du, Z.; Liu, Q.; Chen, B.; Tsang, S.H.; Teo, E.H.T. Ferroelectric BiFeO3 Thin-Film Optical Modulators. Appl. Phys. Lett. 2016, 108, 233502. [Google Scholar] [CrossRef]
  42. Wang, D.Y.; Li, S.; Chan, H.L.W.; Choy, C.L. Optical and Electro-Optic Anisotropy of Epitaxial Ba0.7Sr0.3TiO3 Thin Films. Appl. Phys. Lett. 2010, 96, 061905. [Google Scholar] [CrossRef]
  43. Sando, D.; Yang, Y.; Paillard, C.; Dkhil, B.; Bellaiche, L.; Nagarajan, V. Epitaxial Ferroelectric Oxide Thin Films for Optical Applications. Appl. Phys. Rev. 2018, 5, 041108. [Google Scholar] [CrossRef]
Figure 1. (a) XRD pattern of BST thin film. The inset is the perovskite structure with Ti atom (silver) in center, O atom (red) at face, and Ba/Sr atoms (green) at corner. (b) EDX spectroscopy of BST films. The inset indicates the atomic percentage of each element.
Figure 1. (a) XRD pattern of BST thin film. The inset is the perovskite structure with Ti atom (silver) in center, O atom (red) at face, and Ba/Sr atoms (green) at corner. (b) EDX spectroscopy of BST films. The inset indicates the atomic percentage of each element.
Sensors 22 07183 g001
Figure 2. (a) Survey scanning of BST films. (bf) High resolution XPS of C 1s, Ba 3d, Sr 3d, Ti 2p, and O 1s.
Figure 2. (a) Survey scanning of BST films. (bf) High resolution XPS of C 1s, Ba 3d, Sr 3d, Ti 2p, and O 1s.
Sensors 22 07183 g002
Figure 3. (a) Ferroelectric hysteresis loop of BST thin films. (b) Temperature-dependent dielectric constant and loss of the BST under diverse frequencies.
Figure 3. (a) Ferroelectric hysteresis loop of BST thin films. (b) Temperature-dependent dielectric constant and loss of the BST under diverse frequencies.
Sensors 22 07183 g003
Figure 4. (a) Piezoelectric output voltage on BST at the applied periodic pressure of 10 kPa. (b) Schematic illustration of the piezoelectric effect of sandwiched BST between two electrodes. (c) Stress-triggered output voltage in the range of 10 kPa and 80 kPa. (d) Linear relationship between the performed pressure and the output voltage. The inset indicates the long-term operation stability of this device under the applied pressure of 50 kPa.
Figure 4. (a) Piezoelectric output voltage on BST at the applied periodic pressure of 10 kPa. (b) Schematic illustration of the piezoelectric effect of sandwiched BST between two electrodes. (c) Stress-triggered output voltage in the range of 10 kPa and 80 kPa. (d) Linear relationship between the performed pressure and the output voltage. The inset indicates the long-term operation stability of this device under the applied pressure of 50 kPa.
Sensors 22 07183 g004
Figure 5. (a) Transmittance curve of the resulted BST. The inset reveals the corresponding calculated band gap. (b) Refractive index of BST as a function of wavelength. The inset is the Tan Ψ and Cos Δ from the ellipsometry measurement.
Figure 5. (a) Transmittance curve of the resulted BST. The inset reveals the corresponding calculated band gap. (b) Refractive index of BST as a function of wavelength. The inset is the Tan Ψ and Cos Δ from the ellipsometry measurement.
Sensors 22 07183 g005
Figure 6. (a) Light manipulation as functions of the applied voltage. (b) The relationship between the laser intensity and the applied voltage. The inset is the schematics of experimental setup. (c) Eight possible spontaneous polarization directions in BST along (001). (d) reveals the projected ellipses on (001) plane of BST with and without the applied electric field. Applying an electric field E change the birefringence (Δn) and rotates the optical axis as well.
Figure 6. (a) Light manipulation as functions of the applied voltage. (b) The relationship between the laser intensity and the applied voltage. The inset is the schematics of experimental setup. (c) Eight possible spontaneous polarization directions in BST along (001). (d) reveals the projected ellipses on (001) plane of BST with and without the applied electric field. Applying an electric field E change the birefringence (Δn) and rotates the optical axis as well.
Sensors 22 07183 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, L.; Zhu, M.; Shao, Y.; Zhao, Y.; Wei, C.; Gao, L.; Bao, Y. Smart Sensing Multifunctionalities Based on Barium Strontium Titanate Thin Films. Sensors 2022, 22, 7183. https://doi.org/10.3390/s22197183

AMA Style

Wang L, Zhu M, Shao Y, Zhao Y, Wei C, Gao L, Bao Y. Smart Sensing Multifunctionalities Based on Barium Strontium Titanate Thin Films. Sensors. 2022; 22(19):7183. https://doi.org/10.3390/s22197183

Chicago/Turabian Style

Wang, Linghua, Minmin Zhu, Yong Shao, Yida Zhao, Can Wei, Langfeng Gao, and Yiping Bao. 2022. "Smart Sensing Multifunctionalities Based on Barium Strontium Titanate Thin Films" Sensors 22, no. 19: 7183. https://doi.org/10.3390/s22197183

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