Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 3217 2023-10-24 17:46:21 |
2 Reference format revised. Meta information modification 3217 2023-10-25 03:33:38 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Gaur, A.; Porwal, C.; Boukhris, I.; Chauhan, V.S.; Vaish, R. Multicatalytic Behavior of Ba0.85Ca0.15Ti0.9Zr0.1O3 Ceramic. Encyclopedia. Available online: https://encyclopedia.pub/entry/50755 (accessed on 17 May 2024).
Gaur A, Porwal C, Boukhris I, Chauhan VS, Vaish R. Multicatalytic Behavior of Ba0.85Ca0.15Ti0.9Zr0.1O3 Ceramic. Encyclopedia. Available at: https://encyclopedia.pub/entry/50755. Accessed May 17, 2024.
Gaur, Akshay, Chirag Porwal, Imed Boukhris, Vishal Singh Chauhan, Rahul Vaish. "Multicatalytic Behavior of Ba0.85Ca0.15Ti0.9Zr0.1O3 Ceramic" Encyclopedia, https://encyclopedia.pub/entry/50755 (accessed May 17, 2024).
Gaur, A., Porwal, C., Boukhris, I., Chauhan, V.S., & Vaish, R. (2023, October 24). Multicatalytic Behavior of Ba0.85Ca0.15Ti0.9Zr0.1O3 Ceramic. In Encyclopedia. https://encyclopedia.pub/entry/50755
Gaur, Akshay, et al. "Multicatalytic Behavior of Ba0.85Ca0.15Ti0.9Zr0.1O3 Ceramic." Encyclopedia. Web. 24 October, 2023.
Multicatalytic Behavior of Ba0.85Ca0.15Ti0.9Zr0.1O3 Ceramic
Edit

Ferroelectric materials are known to possess multicatalytic abilities that are nowadays utilized for removing organic pollutants from water via piezocatalysis, photocatalysis, piezo-photocatalysis, and pyrocatalysis processes. The Ba0.85Ca0.15Ti0.9Zr0.1O3 (BCZTO) ceramic is one such ferroelectric composition that has been extensively studied for electrical and electronic applications. Furthermore, the BCZTO ceramic has also shown remarkable multicatalytic performance in water-cleaning applications. 

ferroelectric ceramic multicatalytic activity BCZTO

1. Introduction

Industrial dyes are one of the representative organic contaminants that result in water pollution in terms of both aesthetics and toxicity [1][2]. In addition to endangering humans, contaminated water has a negative impact on marine life as well, which is also a key component of our ecosystem. Catalysis plays a vital role in water-cleaning applications by providing efficient and sustainable methods to remove pollutants, improve water quality, and protect human/environmental health. The development of innovative catalytic materials and processes continues to advance the field of water treatment and contribute to the sustainable management of water resources. Advanced oxidation processes (AOPs) are one of the solutions in water remediation, which comprise a diverse set of oxidation techniques in the aqueous phase that rely on highly reactive species, mainly hydroxyl radicals, to effectively degrade and eliminate targeted pollutants. By harnessing the exceptional reactivity of hydroxyl radicals, these methods initiate intricate oxidation reactions, leading to the breakdown and elimination of various pollutants found in water. While hydroxyl radicals are the primary drivers in AOPs, it is important to acknowledge that other reactive species also contribute to these processes, albeit to a lesser degree [3]. The processes exhibiting AOP characteristics include photocatalysis, piezocatalysis, pyrocatalysis, or a combination of any of these processes.
In the context of water-cleaning applications, ferroelectric materials have emerged as a boon to catalysis processes [4][5]. Ferroelectric materials are a class of materials that exhibit a unique property known as ferroelectricity whereby the spontaneous electric polarization can be reversed by applying an external electric field [6]. Due to this versatility of ferroelectric materials, they are utilized in diverse fields such as energy storage, energy-harvesting systems, actuators, sensors, ultrasound transducers, energy conversion, human health treatment, neuromorphic devices, etc. [7][8][9][10][11][12][13][14].
Focusing on water remediation, the advantageous effect of ferroelectric materials in photocatalysis stems from the presence of a built-in electric field, which effectively facilitates the separation of photogenerated charge carriers [15]. Further, the process involving the generation of polarization induced by mechanical stress, commonly known as piezoelectricity, has been extensively harnessed as a driving force for the catalytic degradation of organic pollutants, even in the absence of light. In recent times, a novel approach to enhancing the degradation efficiency of photocatalysis has emerged, involving the modulation of polarization in ferroelectric materials [15]. This method has shown great promise in boosting photocatalytic performance significantly. The combination of piezoelectricity and photocatalysis holds great potential for efficiently breaking down organic pollutants and holds promise for various environmental remediation applications. It is widely recognized that all ferroelectric materials exhibit pyroelectric properties, enabling them to convert thermal energy into electricity when subjected to temperature fluctuations [16]. Furthermore, the phenomenon of the synergistic effect of possible catalysis processes has been acknowledged as a highly effective method to improve catalytic performance. Additionally, by utilizing ferroelectric switching, it becomes possible to control the chemical properties of catalyst surfaces, leading to a substantial improvement in catalytic efficiency. This opens up new avenues for developing high-performance catalysts by incorporating ferroelectric materials into their design [17].
BaTiO3 (BT) has received the most attention among conventional ferroelectric materials due to its outstanding dielectric, ferroelectric, optical, piezoelectric, and other important characteristics [4][18][19][20][21][22][23][24]. It is well known that the crystal structure of BT stabilizes, depending on temperature, into rhombohedral R3m, orthorhombic Amm2, tetragonal P4mm, and cubic Pm3m phases [25]. Additionally, by adding appropriate impurities or substituting certain crystalline phases in BT, these characteristics and/or crystalline phases can be adjusted and optimized. Ferroelectric ceramics are hardly used in pure chemical form; instead, doping is always used to modify the characteristics for a given application. Donor doping to obtain high piezoelectric coefficients and acceptor doping to achieve low dielectric losses are two examples of how different dopings typically have distinct effects on BT ceramics [26]. Further, equivalent doping at the A-site and B-site is also another approach for tailoring the properties of BT ceramics considering the morphotropic phase boundary (MPB). The simultaneous doping of Ca and Zr in a BT ceramic, making it a (Ba,Ca,Zr,Ti)O3 composition, is one of the most employed compositions for various applications. There are many studies on the (Ba,Ca,Zr,Ti)O3 composition related to dielectric behavior, caloric effects, actuation, sensing, and pyroelectric behavior.

2. Ferroelectric Materials

2.1. Fundamentals of Ferroelectric Materials

There have been numerous developmental milestones since the discovery of ferroelectricity in Rochelle salt about a century ago, which have all been thoroughly examined by Kanzig, Cross, and Newnham, as well as Fousek, and Haertling [27][28][29][30][31][32][33]. The most prevalent characteristic of ferroic materials is the emergence of a domain structure through spontaneously breaking prototypal symmetry, manifesting as hysteresis loops with corresponding conjugate fields. Every ferroelectric material has its distinct hysteresis loop, like a fingerprint. Hysteresis loops allow for the immediate identification of ferroelectricity. 
An increase in the electric field strength causes gradual macroscopic polarisation. The dramatic polarization change near Ec is primarily due to polarization reversal (domain flipping), but at the high-field end, the polarization is saturated and the material behaves as a linear dielectric [6][34]. There is a back-switch in some domains as the electric field intensity decreases, but at zero field the network polarization is non-zero, resulting in remnant polarization (Pr). Since ferroelectrics typically have ferroelastic regions, except for LiNbO3, which only contains 180° of ferroelectric regions, the simultaneous induction of spontaneous strain by an external electric field is also possible [6]. Therefore, a strain–electric field curve, or “butterfly,” can be seen if the strain is measured along with the polarization. The observed hysteresis loops are symmetric in an ideal ferroelectric system, resulting in an equal positive and negative Ec and Pr. Ferroelectric materials, due to their inherent property of spontaneous polarization, have significant applications in various fields such as energy harvesting, energy storage, bio-materials, environmental remediation, and the electronics industry. Several ferroelectric ceramics have been identified for these diverse applications, including BaTiO3 (BT), Pb(Zr,Ti)O3 (PZT), (K,Na)NbO3 (KNN), LiNbO3 (LN), BiFeO3, and Bi4Ti3O12, among others [35][36][37][38][39][40][41][42][43]

2.2. Ba0.85Ca0.15Ti0.9Zr0.1O3 (BCZTO) Ceramics

BCZTO has shown great potential in various emerging energy-harvesting fields due to its numerous advantages. BCZTO is a noteworthy material due to its lead-free composition, which renders it environmentally sustainable [44]. It demonstrates a noteworthy Curie temperature (Tc) of approximately 85 °C, coupled with compelling electrical characteristics [45]. Recent studies have indicated noteworthy advancements in the field of piezoelectric and coupling coefficients of BCZTO. Piezoelectric coefficients (d33) greater than or equal to 630 pC/N and high coupling coefficients (kp) of 0.56 have been observed, which are comparable to those found in the Pb-based materials currently in use [45]. The exceptional values exhibited by BCZTO indicate its potential as a superior alternative. The tunable, electric-field-dependent dielectric properties of BCZTO render it suitable for a wide range of applications. In addition, it has a low dielectric loss of 0.01, which suggests minimal energy dissipation. The favorable amalgamation of characteristics can be ascribed to the concurrent presence of two distinct phases, namely rhombohedral (R) and tetragonal (T), along with an intervening orthorhombic phase [46].
This distinctive arrangement significantly reduces the energy barrier for switching, thereby enhancing the outstanding properties of BCZTO. Considerable research has been carried out to investigate the impact of the composition and MPB on the ferroelectric characteristics of BCZTO [47]. The high piezoelectric properties are achieved at the tricritical point (TCP), which is defined by the simultaneous presence of R, T, and cubic (C) phases at a single point. At the TCP, the energy barrier for the rotation of polarization between the T and R states is low or at its minimum (soft), allowing the direction of polarization to be easily influenced by external pressure or electric fields [48]. This promotes a high level of piezoelectricity and permittivity. Furthermore, it has been reported that lead-free piezomaterial systems contain a triple point where the C–R–T phases coexist. The aforementioned studies have yielded comprehensive insights into the factors that impact the ferroelectric behavior of BCZTO, thereby laying the foundation for further advancements in this promising material.

3. Multicatalytic Ability of Ferroelectric Materials

There are various methods and catalyst materials used for water-cleaning applications, among which titanium dioxide (TiO2) stands out due to its excellent electrochemical properties and stability [49]. However, its efficiency is hampered by significant interior and surface defects, leading to the unwanted recombination of photogenerated electrons and holes during photocatalysis, resulting in reduced energy conversion efficiency [49]. To address these limitations, different techniques have been employed to enhance the charge separation and transport properties in oxide materials, such as surface functionalization, heterojunction formation, and defect adjustment [50][51][52]. Indeed, carbonaceous materials have also been explored as catalyst enhancers for photocatalytic activities. These materials, including activated carbon, carbon nanotubes, and graphene, possess properties that can improve the performance of photocatalysts.
Recently, a novel strategy has emerged involving the use of ferroelectric polarization to control electrochemical processes and modulate the interfacial electronic structure, which has gained considerable attention in photocatalysis [53]. This newfound focus on utilizing ferroelectric and piezoelectric materials in photocatalysis has led to improved charge separation. By combining piezoelectric materials with ferroelectric properties alongside other photocatalysts, the inherent electric field near the piezoelectric material plays a crucial role in facilitating the separation of charges. 

3.1. Photocatalysis Process

In comparison to conventional adsorbents such as active charcoal, metal–organic frameworks, and silica gel, the photocatalytic degradation process offers greater reliability. This is because the recycling of these adsorbents requires harsh conditions to restore their adsorptive capacity, often involving non-environmentally friendly chemicals and solvents that are expensive and lead to secondary pollution [54]. The photocatalytic degradation technique employs an advanced oxidation process. During photocatalysis, highly reactive species are generated, which rapidly attack the organic pollutant molecules and oxidize them in a shorter timeframe. This approach proves more efficient and environmentally friendly, making it a preferred choice over traditional adsorbents.
The process makes use of the energy of illuminating light (visible/UV/solar), whereby the catalyst absorbs energy from the light corresponding to the energy band gap of a catalyst. Indeed, this is the fundamental step involved in all photocatalytic processes. When light with enough energy strikes a semiconductor material, it triggers the excitation of electrons (e) from the valence band (VB) to the conduction band (CB), creating positively charged holes (h+) in the valence band. This photoexcitation process is the starting point for various photocatalytic reactions [37][55]. The photogenerated pairs (e/h+) have the potential to move to the catalyst surface and take part in photocatalytic surface redox reactions before undergoing any recombination process. If the reduction potential of these electrons and holes meets the application requirements, then they can be utilized for diverse photocatalytic reactions.
Based upon the capability of the band energy of the catalyst, the e in the conduction band acts as an oxidizing agent causing adsorbed oxygen (O2) to produce superoxide (.O2) radicals, whereas the h+ in the valence band acts as a strong reducing agent responsible for the production of OH from H2O [56].
Due to their high reactivity, these active species efficiently break down pollutant molecules into smaller fragments, ultimately leading to the complete mineralization of pollutants over time. The degradation reaction takes place on the catalyst’s surface, where the degraded products are released (desorbed) while new pollutant molecules are adsorbed. This cyclic process ensures the completion of the degradation [57].
Utilizing ferroelectric semiconductors offers a viable strategy to enhance photocatalytic activity by introducing a permanent internal polarization. This polarization plays a crucial role in efficiently separating photoexcited carriers. Ferroelectric materials possess distinctive photochemical properties associated with the internal dipole of the material, which stems from the non-centrosymmetric nature of their crystal structure [58]

3.2. Pyrocatalysis Process

The pyroelectric effect is an intriguing phenomenon, which enables catalytic operations to utilize waste heat energy. It has been well-established for a very long time that some crystals respond to temperature variations by forming a spontaneous polarization on their surfaces [59][60]. For many years, pyroelectric materials have been used in conventional applications including temperature sensing and imaging [61]. Microscopic pyroelectric materials have been researched recently for more modern uses, such as surface chemical catalytic processes.
Ferroelectric materials typically exhibit their most significant alteration in polarization and peak pyroelectric characteristics in the vicinity of Tc [62]. As a result, one can anticipate a notably improved catalytic performance in the vicinity of Tc. To prove such a phenomenon, a lead-free BaTi0.89Sn0.11O3 ceramic (Tc ~ 40 °C) was utilized for degrading Rhodamine B (RB) dye in the hot–cold fluctuation range of 25–45 °C (across Tc) and 5–25 °C (below Tc), where the degradation of RB dye was relatively high in the vicinity of its Tc compared to the lowered operating fluctuation temperature [63]. As a fact, when a significant temperature difference is applied to a pyrocatalyst, the charges undergo separation, leading them to engage in redox reactions. 

3.3. Piezocatalysis Process

It is well known that ferroelectric materials are also piezoelectric, thereby providing an immense space for ferroelectric materials to be utilized in the piezocatalysis process. Piezocatalysis is a recently developed, promising technology exhibiting amazing performance in clean energy conversion and water treatment by gathering mechanical energy such as vibration, friction, natural wind, and tides [64]. In 2010, Hong et al. published the first report on the application of piezoelectric materials in chemical catalysis [65]. BaTiO3 microdendrites were used for water splitting, which involves the production of hydrogen and oxygen by splitting water. Further, Hong et al. conducted a study where they reported on the degradation of the organic dye acid orange (AO7) using piezoelectric BaTiO3 microdendrites [66].

3.3.1. Energy Band Gap Theory

This theory says that the piezocatalyst band configuration and electronic states are the primary factors that regulate catalytic activity. In contrast to a perfect insulator, where no mobile charges can move in response to the piezoelectric potential field, the band level tilts linearly across the strained zone, and the oriented accumulation of internal mobile charges can screen the band tilting of piezoelectric semiconductors [67].

3.3.2. Screening Effect Theory

In contrast to the energy band theory, the mechanism of the screening charge effect places more emphasis on the piezopotential’s dominant roles and its associated surface screening behavior in piezoelectric materials. The magnitude of the piezopotential should exactly match the required redox level to be compatible with electrocatalysis and to determine the ability of the piezocatalyst to accomplish a certain chemical reaction.

3.4. Multicatalytic Capability of BCZTO

The phenomenon of dipole randomization after sintering, which is crucial for producing ceramics in bulk form, can hinder catalytic processes. However, this limitation can be overcome by efficiently orienting the dipoles through the application of an electric field. This controlled orientation of dipoles leads to a significant enhancement and improvement in the efficiency of the catalysis process. The ferroelectric BCZTO ceramic exemplifies such a mechanism, where the effective orientation of dipoles contributes to its catalytic prowess.
In a different study, a BCZTO ceramic was employed as a catalyst with a UV light source (365 nm wavelength) for the photocatalysis process, targeting the degradation of RB dye and ciprofloxacin (20 mL and ~10 mg/L concentrated) as representative pollutants [68]. The investigation focused on the impact of conventional poling on the BCZTO ceramic pellet. The results showed that the degradation of RB dye through photocatalysis using UV light was approximately 3.4-fold higher with the poled BCZTO pellet compared to the unpoled BCZTO ceramic pellet. This significant enhancement in photocatalytic efficiency demonstrated the positive effect of poling on the catalytic performance of the BCZTO ceramic. The coupling of piezocatalytic and photocatalytic properties within the ferroelectric material resulted in the increased efficiency of the photocatalytic process. 
The synergistic piezocatalysis and photocatalysis effect is the result of combining the piezocatalytic and photocatalytic processes. When piezoelectric polarization is induced by the ultrasonic wave in the ferroelectric material, electrons and holes are polarized along its spontaneous polarization axis. This creates a p–n junction with a steeper band alignment compared to those activated solely by light irradiation in the photocatalytic process. If polarization is oriented in one direction, then electrons will indeed move in the opposite direction, whereas the holes will follow the opposite path of electrons. These movements of electrons and holes in response to the piezoelectric polarization demonstrate the role of mechanical force in influencing charge transport and separation within the material. This phenomenon is crucial in the synergistic piezo- and photocatalysis effect, as it facilitates enhanced charge separation and contributes to the improved catalytic performance of the material [69].
Moreover, BCZTO ceramics exhibited excellent reusability as a catalyst, showing negligible performance changes even after five cycles of usage in the piezo-photocatalysis process. The effectiveness of poling in the BCZTO ceramic was evident across various processes, including photocatalysis, pyrocatalysis, and piezo-photocatalysis.
Organic pollutant degradation using visible light via the photocatalysis process has become a focus of interest for future prospects. The doping of a transition metal into the parent compound results in reducing the energy band gap of the synthesized ceramics. By incorporating Iron (Fe) into the parent compound, the energy band gap between the conduction and valence bands decreases, enabling the ceramics to absorb visible light. This Fe+3 doping in BCZTO enhances visible light photocatalytic activity by creating defects such as vacancies and narrowing the effective band gap. Similar photocatalysis via visible light has been reported in the literature with Fe doping in BaTiO3 [18]. However, tetragonality in BCZTO diminishes with Fe substitution. In addition, the ferro-para phase transition was diffused following Fe doping, indicating that the long-range interactions between ferroelectric domains and defects caused by Fe were broken. 
It has been reported in the past that adding noble metals (Ag, Au, etc.) to ferroelectric materials can increase the activity in piezocatalysis and cause visible light photocatalysis [70][71]. Carrying forward this approach of enhanced activity in photocatalysis and piezocatalysis processes for degrading organic contaminants, Moolchand et al. loaded Ag onto a BCZTO ceramic and investigated its photo/piezocatalysis performance [72]. The loading of the Ag element onto the BCZTO ceramic was performed by dissolving as-prepared BCZTO ceramic powder and AgNO3 in ethylene glycol solution. The solution was then stirred for around 2 h, after which the solution was repeatedly washed with water and acetone with further exposure to the drying temperature [73]. RB dye was nearly destroyed completely in an aqueous solution in just 40 min utilizing the Ag-loaded BCZTO sample, demonstrating the promising photocatalytic activity of the sample. 
The recovery of catalysts after usage is a challenging task, but one way to facilitate this is by embedding the catalyst in a suitable matrix. This approach becomes even more advantageous when the matrix itself contributes to the catalysis process. One of the novel techniques involves the precipitation of a ceramic phase within a glass matrix, which has been employed as a successful catalyst [37][74]
In the context of its reusability after any catalysis process and being a cost-efficient catalyst, the BCZTO ceramic was mixed with Portland cement in a 50 wt.% proportion and investigated for degrading MB dye via the piezocatalysis process [75]. Cement is one of the most used and common structural materials. Unlike ceramics, which require a specific high-temperature sintering procedure to generate a desired shape, cement-based composites can be cast at ambient temperature in any desired shape and size. 

References

  1. Gupta, V.K. Application of low-cost adsorbents for dye removal—A review. J. Environ. Manag. 2009, 90, 2313–2342.
  2. Crini, G. Non-conventional low-cost adsorbents for dye removal: A review. Bioresour. Technol. 2006, 97, 1061–1085.
  3. Comninellis, C.; Kapalka, A.; Malato, S.; Parsons, S.A.; Poulios, I.; Mantzavinos, D. Advanced oxidation processes for water treatment: Advances and trends for R&D. J. Chem. Technol. Biotechnol. Int. Res. Process. Environ. Clean Technol. 2008, 83, 769–776.
  4. Gaur, A.; Dubey, S.; Elqahtani, Z.M.; Ahmed, S.B.; Al-Buriahi, M.S.A.; Vaish, R.; Chauhan, V.S. Effect of Poling on Multicatalytic Performance of 0.5Ba(Zr0.2Ti0.8)O3-0.5(Ba0.7Sr0.3)TiO3 Ferroelectric Ceramic for Dye Degradation. Materials 2022, 15, 8217.
  5. Gaur, A.; Chauhan, V.S.; Vaish, R. Planetary ball milling induced piezocatalysis for dye degradation using BaTiO3 ceramics. Environ. Sci. Adv. 2023, 2, 462–472.
  6. Jin, L.; Li, F.; Zhang, S. Decoding the fingerprint of ferroelectric loops: Comprehension of the material properties and structures. J. Am. Ceram. Soc. 2014, 97, 1–27.
  7. Wei, X.; Domingo, N.; Sun, Y.; Balke, N.; Dunin-Borkowski, R.E.; Mayer, J. Progress on emerging ferroelectric materials for energy harvesting, storage and conversion. Adv. Energy Mater. 2022, 12, 2201199.
  8. Mikolajick, T.; Park, M.H.; Begon-Lours, L.; Slesazeck, S. From Ferroelectric Material Optimization to Neuromorphic Devices. Adv. Mater. 2022, 2206042.
  9. Butler, K.T.; Frost, J.M.; Walsh, A. Ferroelectric materials for solar energy conversion: Photoferroics revisited. Energy Environ. Sci. 2015, 8, 838–848.
  10. Yuan, Y.; Xiao, Z.; Yang, B.; Huang, J. Arising applications of ferroelectric materials in photovoltaic devices. J. Mater. Chem. A 2014, 2, 6027–6041.
  11. Bowen, C.R.; Kim, H.A.; Weaver, P.M.; Dunn, S. Piezoelectric and ferroelectric materials and structures for energy harvesting applications. Energy Environ. Sci. 2014, 7, 25–44.
  12. Xu, Y. Ferroelectric Materials and Their Applications; Elsevier: Amsterdam, The Netherlands, 2013; ISBN 1483290956.
  13. Tagantsev, A.K.; Sherman, V.O.; Astafiev, K.F.; Venkatesh, J.; Setter, N. Ferroelectric materials for microwave tunable applications. J. Electroceram. 2003, 11, 5–66.
  14. Yang, D.; Zhang, R.; Gai, S.; Yang, P. The fundamental and application of piezoelectric materials for tumor therapy: Recent advances and outlook. Mater. Horiz. 2023, 10, 1140–1184.
  15. Wu, J.; Xu, Q.; Lin, E.; Yuan, B.; Qin, N.; Thatikonda, S.K.; Bao, D. Insights into the role of ferroelectric polarization in piezocatalysis of nanocrystalline BaTiO3. ACS Appl. Mater. Interfaces 2018, 10, 17842–17849.
  16. Chen, J.; Luo, W.; Yu, S.; Yang, X.; Wu, Z.; Zhang, H.; Gao, J.; Mai, Y.-W.; Li, Y.; Jia, Y. Synergistic effect of photocatalysis and pyrocatalysis of pyroelectric ZnSnO3 nanoparticles for dye degradation. Ceram. Int. 2020, 46, 9786–9793.
  17. Ding, W.; Lu, J.; Tang, X.; Kou, L.; Liu, L. Ferroelectric Materials and Their Applications in Activation of Small Molecules. ACS Omega 2023, 8, 6164–6174.
  18. Gaur, A.; Sharma, M.; Chauhan, V.S.; Vaish, R. Solar/visible light photocatalytic dye degradation using BaTi1-xFexO3 ceramics. J. Am. Ceram. Soc. 2022, 105, 5140–5150.
  19. Shih, W.Y.; Shih, W.-H.; Aksay, I.A. Size dependence of the ferroelectric transition of small BaTiO3 particles: Effect of depolarization. Phys. Rev. B 1994, 50, 15575.
  20. Gao, J.; Liu, Y.; Wang, Y.; Hu, X.; Yan, W.; Ke, X.; Zhong, L.; He, Y.; Ren, X. Designing high dielectric permittivity material in barium titanate. J. Phys. Chem. C 2017, 121, 13106–13113.
  21. Woldu, T.; Raneesh, B.; Sreekanth, P.; Reddy, M.V.R.; Philip, R.; Kalarikkal, N. Size dependent nonlinear optical absorption in BaTiO3 nanoparticles. Chem. Phys. Lett. 2015, 625, 58–63.
  22. Wang, Z.; Suryavanshi, A.P.; Yu, M.-F. Ferroelectric and piezoelectric behaviors of individual single crystalline BaTiO3 nanowire under direct axial electric biasing. Appl. Phys. Lett. 2006, 89, 82903.
  23. Von Hippel, A. Ferroelectricity, domain structure, and phase transitions of barium titanate. Rev. Mod. Phys. 1950, 22, 221.
  24. Everhardt, A.S.; Denneulin, T.; Grünebohm, A.; Shao, Y.-T.; Ondrejkovic, P.; Zhou, S.; Domingo, N.; Catalan, G.; Hlinka, J.; Zuo, J.-M. Temperature-independent giant dielectric response in transitional BaTiO3 thin films. Appl. Phys. Rev. 2020, 7, 11402.
  25. Kwei, G.H.; Lawson, A.C.; Billinge, S.J.L.; Cheong, S.W. Structures of the ferroelectric phases of barium titanate. J. Phys. Chem 1993, 97, 2368–2377.
  26. Liu, W.; Cheng, L.; Li, S. Prospective of (BaCa)(ZrTi)O3 lead-free piezoelectric ceramics. Crystals 2019, 9, 179.
  27. Haertling, G.H. Ferroelectric ceramics: History and technology. J. Am. Ceram. Soc. 1999, 82, 797–818.
  28. Valasek, J. Piezo-electric and allied phenomena in Rochelle salt. Phys. Rev. 1921, 17, 475.
  29. Valasek, J. Piezo-electric activity of Rochelle salt under various conditions. Phys. Rev. 1922, 19, 478.
  30. Valasek, J. Properties of Rochelle salt related to the piezo-electric effect. Phys. Rev. 1922, 20, 639.
  31. Kanzig, W. History of ferroelectricity 1938–1955. Ferroelectrics 1987, 74, 285–291.
  32. Cross, L.E.; Newnham, R.E. History of ferroelectrics. Ceram. Civiliz. 1987, 3, 289–305.
  33. Fousek, J. Ferroelectricity: Remarks on historical aspects and present trends. Ferroelectrics 1991, 113, 3–20.
  34. Hu, Z.; Zhang, H.; Zhu, Z.; Zeng, J.; Zhang, D.; Eriksson, M.; Viola, G.; Yan, H. Deciphering the peculiar hysteresis loops of 0.05 Pb (Mn1/3Sb2/3) O3–0.95 Pb (Zr0.52Ti0.48) O3 piezoelectric ceramics. Acta Mater. 2023, 244, 118563.
  35. Kong, Y.; Liu, S.; Zhao, Y.; Liu, H.; Chen, S.; Xu, J. Highly optical damage resistant crystal: Zirconium-oxide-doped lithium niobate. Appl. Phys. Lett. 2007, 91, 081908.
  36. Stock, M.; Dunn, S. Influence of the ferroelectric nature of lithium niobate to drive photocatalytic dye decolorization under artificial solar light. J. Phys. Chem. C 2012, 116, 20854–20859.
  37. Gaur, A.; Sharma, M.; Chauhan, V.S.; Vaish, R. Visible light photocatalytic activity in BiFeO3 glass-ceramics. Mater. Chem. Phys. 2023, 303, 127710.
  38. Wu, J.; Qin, N.; Lin, E.; Yuan, B.; Kang, Z.; Bao, D. Synthesis of Bi4Ti3O12 decussated nanoplates with enhanced piezocatalytic activity. Nanoscale 2019, 11, 21128–21136.
  39. Yuan, L.; Fan, W.; Yang, X.; Ge, S.; Xia, C.; Foong, S.Y.; Liew, R.K.; Wang, S.; Van Le, Q.; Lam, S.S. Piezoelectric PAN/BaTiO3 nanofiber membranes sensor for structural health monitoring of real-time damage detection in composite. Compos. Commun. 2021, 25, 100680.
  40. Dai, Z.; Li, D.; Zhou, Z.; Zhou, S.; Liu, W.; Liu, J.; Wang, X.; Ren, X. A strategy for high performance of energy storage and transparency in KNN-based ferroelectric ceramics. Chem. Eng. J. 2022, 427, 131959.
  41. Kang, M.-G.; Jung, W.-S.; Kang, C.-Y.; Yoon, S.-J. Recent progress on PZT based piezoelectric energy harvesting technologies. Actuators 2016, 5, 5.
  42. Roscow, J.I.; Lewis, R.W.C.; Taylor, J.; Bowen, C.R. Modelling and fabrication of porous sandwich layer barium titanate with improved piezoelectric energy harvesting figures of merit. Acta Mater. 2017, 128, 207–217.
  43. Liu, D.; Jin, C.; Zhang, Y.; He, Y.; Wang, F. Integrated piezo-photocatalysis of electrospun Bi4Ti3O12 nanostructures by bi-harvesting visible light and ultrasonic energies. Ceram. Int. 2021, 47, 7692–7699.
  44. Ji, W.; Fang, B.; Lu, X.; Zhang, S.; Yuan, N.; Ding, J. Tailoring structure and performance of BCZT ceramics prepared via hydrothermal method. Phys. B Condens. Matter 2019, 567, 65–78.
  45. Wang, P.; Li, Y.; Lu, Y. Enhanced piezoelectric properties of (Ba0.85Ca0.15)(Ti0.9Zr0.1O3) O3 lead-free ceramics by optimizing calcination and sintering temperature. J. Eur. Ceram. Soc. 2011, 31, 2005–2012.
  46. Turygin, A.P.; Neradovskiy, M.M.; Naumova, N.A.; Zayats, D.V.; Coondoo, I.; Kholkin, A.L.; Shur, V.Y. Domain structures and local switching in lead-free piezoceramics Ba0.85Ca0.15Ti0.90Zr0.10O3. J. Appl. Phys. 2015, 118, 72002.
  47. Yao, S.; Ren, W.; Ji, H.; Wu, X.; Shi, P.; Xue, D.; Ren, X.; Ye, Z.-G. High pyroelectricity in lead-free 0.5Ba(Zr0.2Ti0.8) O3–0.5 (Ba0.7Ca0.3)TiO3 ceramics. J. Phys. D Appl. Phys. 2012, 45, 195301.
  48. Sahoo, B.; Thejas, T.S.; Politova, E.D.; Panda, P.K. Effect of dopants on electrical properties of BCT-BZT lead free piezo ceramics: A review. Ferroelectrics 2021, 582, 46–62.
  49. Li, Y.; Li, J.; Yang, W.; Wang, X. Implementation of ferroelectric materials in photocatalytic and photoelectrochemical water splitting. Nanoscale Horiz. 2020, 5, 1174–1187.
  50. Cao, F.; Xiong, J.; Wu, F.; Liu, Q.; Shi, Z.; Yu, Y.; Wang, X.; Li, L. Enhanced photoelectrochemical performance from rationally designed anatase/rutile TiO2 heterostructures. ACS Appl. Mater. Interfaces 2016, 8, 12239–12245.
  51. Wang, X.; Li, Z.; Shi, J.; Yu, Y. One-dimensional titanium dioxide nanomaterials: Nanowires, nanorods, and nanobelts. Chem. Rev. 2014, 114, 9346–9384.
  52. Yu, Y.; Zhang, Z.; Yin, X.; Kvit, A.; Liao, Q.; Kang, Z.; Yan, X.; Zhang, Y.; Wang, X. Enhanced photoelectrochemical efficiency and stability using a conformal TiO2 film on a black silicon photoanode. Nat. Energy 2017, 2, 17045.
  53. Liang, Z.; Yan, C.-F.; Rtimi, S.; Bandara, J. Piezoelectric materials for catalytic/photocatalytic removal of pollutants: Recent advances and outlook. Appl. Catal. B Environ. 2019, 241, 256–269.
  54. Ali, I. New generation adsorbents for water treatment. Chem. Rev. 2012, 112, 5073–5091.
  55. Porwal, C.; Verma, S.; Chauhan, V.S.; Vaish, R. Bismuth zinc borate-Polyacrylonitrile nanofibers for photo-piezocatalysis. J. Ind. Eng. Chem. 2023, 124, 358–367.
  56. Teh, C.M.; Mohamed, A.R. Roles of titanium dioxide and ion-doped titanium dioxide on photocatalytic degradation of organic pollutants (phenolic compounds and dyes) in aqueous solutions: A review. J. Alloys Compd. 2011, 509, 1648–1660.
  57. Kumar, A.; Kumar, A.; Krishnan, V. Perovskite oxide based materials for energy and environment-oriented photocatalysis. ACS Catal. 2020, 10, 10253–10315.
  58. Khan, M.A.; Nadeem, M.A.; Idriss, H. Ferroelectric polarization effect on surface chemistry and photo-catalytic activity: A review. Surf. Sci. Rep. 2016, 71, 1–31.
  59. Zhang, Y.; Phuong, P.T.T.; Roake, E.; Khanbareh, H.; Wang, Y.; Dunn, S.; Bowen, C. Thermal energy harvesting using pyroelectric-electrochemical coupling in ferroelectric materials. Joule 2020, 4, 301–309.
  60. Marvan, M.; Fousek, J. Pyroelectricity and compositionally graded ferroelectrics: A brief review of our theoretical approach. Phase Transit. 2006, 79, 485–492.
  61. Bowen, C.R.; Taylor, J.; LeBoulbar, E.; Zabek, D.; Chauhan, A.; Vaish, R. Pyroelectric materials and devices for energy harvesting applications. Energy Environ. Sci. 2014, 7, 3836–3856.
  62. Xu, X.; Xiao, L.; Jia, Y.; Wu, Z.; Wang, F.; Wang, Y.; Haugen, N.O.; Huang, H. Pyro-catalytic hydrogen evolution by Ba 0.7 Sr 0.3 TiO3 nanoparticles: Harvesting cold–hot alternation energy near room-temperature. Energy Environ. Sci. 2018, 11, 2198–2207.
  63. Zhao, Q.; Xiao, H.; Huangfu, G.; Zheng, Z.; Wang, J.; Wang, F.; Guo, Y. Highly-efficient piezocatalytic performance of nanocrystalline BaTi0.89Sn0.11O3 catalyst with Tc near room temperature. Nano Energy 2021, 85, 106028.
  64. Liu, D.; Zhang, J.; Tan, L.; Jin, C.; Li, M.; Chen, B.; Zhang, G.; Zhang, Y.; Wang, F. Enhanced piezocatalytic hydrogen evolution performance of bismuth vanadate by the synergistic effect of facet engineering and cocatalyst engineering. J. Colloid Interface Sci. 2023, 646, 159–166.
  65. Hong, K.-S.; Xu, H.; Konishi, H.; Li, X. Direct water splitting through vibrating piezoelectric microfibers in water. J. Phys. Chem. Lett. 2010, 1, 997–1002.
  66. Hong, K.-S.; Xu, H.; Konishi, H.; Li, X. Piezoelectrochemical effect: A new mechanism for azo dye decolorization in aqueous solution through vibrating piezoelectric microfibers. J. Phys. Chem. C 2012, 116, 13045–13051.
  67. Starr, M.B.; Wang, X. Coupling of piezoelectric effect with electrochemical processes. Nano Energy 2015, 14, 296–311.
  68. Sharma, M.; Singh, V.P.; Kumar, S.; Vaish, R. Multicatalytic behavior of Ba0.85Ca0.15Ti0.9Zr0.1O3 ceramics for pharmaceutical/dye/bacterial treatments. J. Appl. Phys. 2020, 127, 135103.
  69. Liu, Y.-L.; Wu, J.M. Synergistically catalytic activities of BiFeO3/TiO2 core-shell nanocomposites for degradation of organic dye molecule through piezophototronic effect. Nano Energy 2019, 56, 74–81.
  70. Xu, S.; Liu, Z.; Zhang, M.; Guo, L. Piezotronics enhanced photocatalytic activities of Ag-BaTiO3 plasmonic photocatalysts. J. Alloys Compd. 2019, 801, 483–488.
  71. Nithya, P.M.; Devi, L.G. Effect of surface Ag metallization on the photocatalytic properties of BaTiO3: Surface plasmon effect and variation in the Schottky barrier height. Surf. Interfaces 2019, 15, 205–215.
  72. Sharma, M.; Singh, G.; Vaish, R. Ag-nanoparticles-loaded Ba0.85Ca0.15Ti0.9Zr0.1O3 for multicatalytic dye degradation. Nanotechnology 2021, 32, 145716.
  73. Luo, S.; Yu, S.; Fang, F.; Lai, M.; Sun, R.; Wong, C.-P. Investigating the mechanism of catalytic reduction of silver nitrate on the surface of barium titanate at room temperature: Oxygen vacancies play a key role. RSC Adv. 2015, 5, 3377–3380.
  74. Gaur, A.; Sharma, M.; Chauhan, V.S.; Vaish, R. BaTiO3 crystallized glass-ceramic for water cleaning application via piezocatalysis. Nano-Struct. Nano-Objects 2023, 35, 101005.
  75. Sharma, M.; Vaish, R. Vibration energy harvesting for degradation of dye and bacterial cells using cement-based Ba0.85Ca0.15Zr0.1Ti0.90O3 composites. Mater. Today Commun. 2020, 25, 101592.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , ,
View Times: 224
Revisions: 2 times (View History)
Update Date: 25 Oct 2023
1000/1000