Next Article in Journal
Nanoscale Printing of Indium-Tin-Oxide by Femtosecond Laser Pulses
Next Article in Special Issue
Effect of Nano-Sized γ′ Phase on the Ultrasonic and Mechanical Properties of Ni-Based Superalloy
Previous Article in Journal
Analysis of Formation Mechanisms of Sugar-Derived Dense Carbons via Hydrogel Carbonization Method
Previous Article in Special Issue
Fabrication and Performance of UV Photodetector of ZnO Nanorods Decorated with Al Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Lead-Free Bi0.5Na0.5TiO3 Ferroelectric Nanomaterials for Pyro-Catalytic Dye Pollutant Removal under Cold-Hot Alternation

1
Xi’an Key Laboratory of Textile Chemical Engineering Auxiliaries, School of Environmental and Chemical Engineering, Xi’an Polytechnic University, Xi’an 710600, China
2
School of Science, Xi’an University of Posts and Telecommunications, Xi’an 710121, China
3
College of Materials and Chemistry, China Jiliang University, Hangzhou 310018, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(22), 4091; https://doi.org/10.3390/nano12224091
Submission received: 21 October 2022 / Revised: 17 November 2022 / Accepted: 19 November 2022 / Published: 21 November 2022

Abstract

:
In this work, explicitly pyro-catalytic performance is observed in sol-gel-synthesized ferroelectric Bi0.5Na0.5TiO3 lead-free nanomaterials, and its application for dye wastewater purification is also actualized under temperature fluctuations varying from 23 °C to 63 °C. The decomposition ratios of the pyro-catalytic Bi0.5Na0.5TiO3 nanomaterials on Rhodamine B, methyl blue and methyl orange can reach 96.75%, 98.35% and 19.97%, respectively. In the pyro-catalytic process, the probed active species such as hydroxyl radicals, superoxide radicals and holes play an extremely important role in decomposing dye molecules. The ferroelectric Bi0.5Na0.5TiO3 lead-free nanomaterials will have an excellent prospect for dye wastewater purification due to its explicit pyro-catalysis.

1. Introduction

In recent years, with the increasing demand for color in textiles and printing, the textile industry has developed rapidly [1,2,3,4,5]. Numerous organic dyes sneak into the water, causing serious water pollution and arousing people’s concern [6,7]. To change this situation, numerous methods have continuously been discovered for purifying dye wastewater [8,9,10,11,12,13], in which photocatalysis technology, as an advanced oxidation method, has caused a great sensation in recent decades [14]. In the photocatalytic process, photo-induced electrons and holes will participate in the chemical reactions to form active species with strong oxidation abilities such as hydroxyl radicals (•OH) and superoxide anions ( O 2 ) for dye wastewater purification [15,16]. Some photocatalysts with excellent dye decomposition performance have been reported, such as ZnO [17] and TiO2 [18]. However, the commercial photocatalyst is restricted to some extent, mainly owing to the narrow response range of the visible light wavelength and the low utilization efficiency of solar energy [19]. Therefore, it is essential to seek new catalytic technology for dye wastewater purification.
Apart from sunlight, temperature fluctuations are also an extremely common phenomenon and deemed a readily available energy source [20]. Pyroelectric materials possess the ability to convert thermal energy into electrical energy in the case of temperature fluctuations [21,22,23,24,25,26,27]. Theoretically speaking, it is possible to use the pyroelectrically induced negative and positive electric charges to generate strongly active species and catalytically decompose dye molecules [28,29,30,31,32,33,34,35,36,37]. Recently, some reported materials such as ZnO [38], BaTiO3 [39] and NaNbO3 [40] can exhibit highly pyro-catalytic activity for purifying dye wastewater, which means that a newly catalytic method via utilizing the natural energy of cold-hot fluctuation can be developed [41,42,43,44,45,46,47,48,49,50].
In the ferroelectric families, lead-free bismuth sodium titanate (Bi1-xNaxTiO3), a sort of perovskite structure, possesses many different oxidation layers and a superb pyroelectric property near the morphotropic phase boundary (MPB) (x = 0.5) [51,52]. It has been reported that Bi0.5Na0.5TiO3 (BNT) possesses the optimal pyroelectric response with a relatively high Curie temperature of 320 °C [53,54], a pyroelectric coefficient (p) of 2.5 × 10−4 C · m−2 · K−1 and no toxicity to the environment [55,56]. In addition, nanoscale catalysts often possess excellent catalytic performance, mainly owing to their large surface areas [57]. Therefore, nanoscale BNT is an ideal candidate material to be used to investigate the pyro-catalysis for purifying dye wastewater. Up to now, there has been rare reporting on the pyro-catalytic dye decomposition using the pyroelectric BNT lead-free nanomaterials near the morphotropic phase boundary region as the candidate catalyst.
In this study, strongly pyro-catalytic activity for purifying dye wastewater is achieved in sol-gel-synthesized lead-free ferroelectric BNT nanomaterials under cold-hot fluctuation varying from 23 °C to 63 °C. The pyro-catalytic decomposition ratios of BNT nano-catalysts on Rhodamine B (RhB) dye, methyl blue (MB) dye and methyl orange (MO) dye can reach 96.75%, 98.35% and 19.97%, respectively.

2. Materials and Methods

2.1. Synthesis of BNT

The BNT nanocatalysts were synthesized using a sol-gel method. All the raw materials were of analytical grades. Bismuth nitrate pentahydrate (Bi(NO3)3 · 5H2O) and sodium acetate (CH3COONa) were dissolved in an acetic acid (CH3COOH) solution. After stirring for 3 h, a stoichiometric tetrabutyl titanate (Ti(C4H9O)4) solution was added to form a complex sol. Subsequently, the deionized water was appended to the mixed solution (the molar ratio of deionized water to titanium solution was 50:1). Ethylene glycol ether (CH3OCH2CH2OH) was added to stabilize the Ti(C4H9O)4. Then, the PH value of the solution was adjusted to two. The mixture was heated at 55 °C to form a clear and stable precursor. The synthesized precursor was transferred to a 120 °C oven to dry. The dried gel was heated and kept at 650 °C for 1 h.

2.2. Characterization

The crystalline structures of the synthesized BNT were characterized on an X-ray diffraction (XRD, Philips PW3040/60, Eindhoven, The Netherlands) using monochromatic Cu Kα radiation (λ = 1.5406 Å, 2θ = 20°–80°). The surface morphology was characterized by scanning electron microscopy (SEM, Phenom ProX desktop, Alemlo, The Netherlands). The morphology of the BNT was further observed with a transmission electron microscope (TEM, JEOL 2010F, Japan) with an acceleration voltage of 200 kV. The elements in the BNT were analyzed through energy-dispersive X-ray spectroscopy (EDS, Phenom proX, Alemlo, The Netherlands). The ferroelectric hysteresis loop was measured at room temperature through employing a ferroelectric analyzer (RTI-Multiferroic-4 kV, Radiant, Albuquerque, NM, USA) at 100 Hz. To characterize the chemical element composition in the BNT sample, X-ray photoelectron spectroscopy (XPS, ESCALAB 250Xi, Waltham, MA, USA) analysis was performed through using a Thermo Fisher Scientific Ka in wide scan survey mode and a high-energy resolution with Al Ka. All the binding energies were calibrated through using a C 1s peak at 284.8 eV of the surface adventitious carbon as a reference.

2.3. Pyro-Catalysis Performance Test

These solutions of RhB, MO and MB were selected for testing the pyro-catalytic activities. First, 50 mL of RhB dye (5 mg/L), MO dye (5 mg/L) and MB dye (5 mg/L) solutions were put into three 100-mL glass beakers, respectively. The BNT nano-powders were evenly dispersed in the three types of dye solutions. Before the cold-hot fluctuation, each dye solution with BNT was mechanically stirred for 1 h to achieve the adsorption-desorption equilibrium between the dye solution and catalyst. Here, a water bath was used to achieve uniform cold-hot fluctuation varying from 23 °C to 63 °C in the dark to avoid photodecomposition. Again, every 10 cold-hot cycles, a 2-mL dye solution sample was collected and centrifuged to separate the dye solution and catalyst. The absorption spectra of these dye solutions were recorded through a Hitachi U-3900 (Japan) UV–Vis spectrophotometer.

2.4. Detection of Pyro-Catalytic Active Species

The trapping experiments were designed to make certain the active species during the pyro-catalytic process. Iso-propanol (IPA, •OH scavenger) [58], p-benzoquinone (BQ, O 2 scavenger) and ethylenediaminetetraacetate (EDTA, a quencher of holes) were selected as inhibitors [59,60]. The RhB was selected as the experimental dye for trapping of the active species. In the process of the experiment, 2 mL of the dye solution was collected every 10 cold-hot cycles for UV–Vis spectrophotometer measurement.

3. Results and Discussion

Figure 1 shows the microstructure characterization of BNT. Figure 1a depicts the XRD of the synthesized BNT nanopowders, in which the characteristic peaks of 22.8°, 32.6°, 40.2°, 46.7°, 58.2°, 68.4° and 77.8° correspond to the characteristic planes (110), (101), (012), (220), ( 1 ¯ 23 ) , ( 2 ¯ 42 ) and ( 2 ¯ 43 ) of monoclinic phase BNT (PDF card number 46-0001), respectively. All characteristic peaks could be observed in the diffractogram, confirming the high purity of the BNT. The TEM image of BNT can be presented through the inset of Figure 1a, which shows that the BNT powders were on the nanometer scale of ~100 nm with a polygonal shape. The EDX image of BNT in Figure 1b clearly confirms the presence of Na, O, Ti and Bi, where the probed Au and Al resulted from the sputtered Au coating on the powder sample and sample holder, respectively. The SEM image is illustrated in the inset of Figure 1b, in which the BNT powder samples show smooth edges and flat surfaces with a well-defined polygonal morphology.
The survey XPS spectra of BNT is illustrated in Figure 2, in which these peaks of Bi 4f, Na 1s, Ti 2p and O 1s for BNT can be intuitively observed. These unmarked peaks correspond with these elemental oxidation states.
Figure 3 further confirms results of the local magnification of XPS. In Figure 3a, the Bi 4f spectrum originated from two contributions, 4f7/2 and 4f5/2, located at 160.6 eV and 165.9 eV, respectively. The Bi 4f spectrum clearly proves the presence of two chemical environments for Bi atoms. The two strong peaks are allocated to the Bi 4f7/2 and Bi 4f5/2 peaks of Bi3+ [61], respectively. In Figure 3b, The Na 1s peak was detected at 1072.93 eV. The peak corresponds to the Na+ cation [62]. In Figure 3c, the Ti 2p spectrum also originated from two contributions, Ti 2p3/2 and Ti 2p1/2, located at 459.24 eV and 465.34 eV, respectively [63]. Figure 3d shows the binding energy value of O 1s. The O 1s peak was detected at 531.32 eV.
The ferroelectric loop of BNT can be seen in Figure 4, in which the well-saturated loop indicates the ferroelectric characterization of BNT. The inset of Figure 4 demonstrates the schematic diagram of the circuit for the ferroelectric measurement, with the BNT laid on an insulated glass substrate and the two electrodes made of Au. The pyroelectric properties of the BNT materials were expected, since all the ferroelectric materials were pyroelectric [64].
Figure 5 shows the adsorption spectra of the RhB dye solution with BNT under temperature fluctuations varying from 23 °C to 63 °C. The inset of Figure 5 shows the designed temperature fluctuation curve (from 23 °C to 63 °C) for cold-hot cycles during the pyro-catalytic process. Every dye had its own maximum absorption peak intensity. The absorption peak intensity of the RhB dye was located at ~554 nm. With the increase in cold-hot cycles, the absorption peak height of the RhB dye solution gradually decreased, indicating the excellent pyro-catalytic activities of BNT.
Figure 6 depicts the mechanism schematic diagram of pyro-catalysis under the cold-hot fluctuation. When the temperature is altered, these positive and negative charges will be generated on the surfaces of the pyroelectric BNT catalysts based on Equation (1) [65,66]:
BNT Δ T BNT + q + + q
The extra electric charges be equipoised by the charge compensation of O 2 and OH in a water solution. These negative charges ( q ) will produce the activated superoxide radicals (• O 2 ) based on Equation (2):
O 2 + q     O 2
Similarly, the positive charges ( q + ) will produce the hydroxyl radicals (•OH) based on Equation (3):
OH + q +   OH
Finally, these active species decompose stubborn dye molecules based on Equation (4):
OH   or   O 2 + Dyes Dye   decomposition
Figure 7 presents the results of the decomposition ratio of RhB dye under different conditions and the surface radical trapping experiments, in which the decomposition ratio (D) of the RhB dye can be calculated as D = (1 − C/C0), where C0 (C0 = 5 mg/L) and C are the initial concentration and the actual concentration of the dye solution during the pyro-catalytic process, respectively. In Figure 7, the decomposition of the RhB dye was significantly inhibited by IPA, and the addition of EDTA and BQ was also heavily inhibited, Furthermore, the color change of the RhB dye solution with BNT is shown in the inset of Figure 7, in which the color of the RhB dye solution became transparent after 85 cold-hot cycles. These results indicate that RhB dye is difficult to self-decompose under temperature fluctuations, the decomposition behavior of RhB dye results from the pyro-catalysis, and the extreme importance of •OH is shown, similarly indicating that O 2 and holes are the main active species in the pyro-catalytic process.
Figure 8 shows the decomposition results of the MB dye and MO dye, in which the adsorption spectra of the MB and MO solutions correspond to Figure 8a,b, respectively. The absorption peak intensities of the MB dye and MO dye are located at ~664 nm and ~485 nm, respectively. With the increase in cold-hot cycles, the absorption peak heights of the two dyes gradually decreased. Furthermore, the color changes of the MB dye and MO dye are exhibited in the inset of Figure 8a,b, respectively. The color in the MB dye solution became transparent after 35 cold-hot cycles, while the color in the MO dye solution changed to almost nothing after 85 cold-hot cycles.
The pyro-catalytic activities of BNT for different dye solutions are reflected through the decomposition ratios in Figure 9. The decomposition ratios of the RhB dye and MO dye solution reached 96.75% and 19.97%, respectively, after undergoing 95 cold-hot cycles, while the decomposition ratio of the MB solution was 98.35% after undergoing 35 cold-hot cycles. These different decomposition activities may be associated with their organic molecular chain structures [67], which cause different decomposition mechanisms. In theory, decomposition reactions begin to occur between a dye molecule and a strong oxidation radical [68]. The decomposition process was accelerated, owing to the de-ethylating reaction in the RhB dye solution, causing high pyro-catalytic activity [69,70]. During the radical reactions, the sulfur located on the aromatic links in the MB dye molecules was more reactive than the sulfonyl group in the p-benzene ring of the MO dye molecules. Additionally, the molecules of MB and RhB are much larger in size than those of the MO dye, which is conducive to being absorbed and broken [71]. As shown in Figure 9, the decomposition ratio of the MO dye was the lowest.
In addition to the pyro-catalytic activity exhibited in this work, Li et al. have reported that the BNT nano-catalyst also possesses remarkable photocatalytic performance, with a ~91.6% decomposition ratio for MO dye [72]. Recently, a practical method named light-thermal synergy to further enhance catalytic decomposition ratio was reported [73]. D. Jiang et al. have reported that the catalytic dye decomposition ratio through thermo-or photo-bi-catalysis synergy is further enhanced compared with that of pyro-catalysis or photocatalysis itself [74]. Therefore, the amelioration of catalytic activities in BNT may be expectable, through the synergy of thermo- or photo-bi-catalysis in the future [75,76,77].

4. Conclusions or Summary

In summary, a remarkable pyro-catalytic phenomenon for dye wastewater purification was observed in sol-gel-synthesized BNT ferroelectric lead-free nanomaterials under cold-hot fluctuation varying from 23 °C to 63 °C. The pyro-catalytic decomposition ratios of the BNT nanomaterials on RhB dye, MB dye and MO dye were 96.75%, 98.35% and 19.97%, respectively. The catalysis reaction on the BNT catalysts’ surfaces under the environmental temperature fluctuation was excited by these pyroelectrically induced active species with strong oxidation abilities. This promising pyro-catalysis in lead-free BNT ferroelectric nanomaterials has potential in purifying dye wastewater utilizing cold-hot fluctuations.

Author Contributions

Conceptualization, Z.W., S.H. and Y.J.; methodology, Z.W., S.W. and S.H.; validation, Z.W., S.W. and X.S.; formal analysis, Z.W., X.X. and Z.C.; investigation, D.G., Y.Z. and X.X.; resources, Z.C., X.S. and Y.J.; data curation, Z.W.; writing—original draft preparation, Z.W. and S.W.; writing—review and editing, Z.W.; visualization, Z.W., D.G., Y.Z. and Y.J.; supervision, Z.W., S.H. and Y.J.; project administration, Z.W. and Y.J.; funding acquisition, Z.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (grant numbers 51872264 and 22179108), Key Research and Development Projects of Shaanxi Province (grant number 2020GXLH-Z-032), Shaanxi Provincial Natural Science Foundation of China (grant number 2020JM-579), Doctoral Research Start-up Fund project of Xi’an Polytechnic University (grant number 107020589 and 107020511), Shaanxi Provincial High-Level Talents Introduction Project (Youth Talent Fund), the Key Research and Development Program of Ningbo (2022Z178), and China Construction Technology Research and Development Project (CSCEC-2021-Z-5-02).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lin, H.; Wu, Z.; Jia, Y.; Li, W.; Zheng, R.K.; Luo, H.S. Piezoelectrically induced mechano-catalytic effect for degradation of dye wastewater through vibrating Pb(Zr0.52Ti0.48)O3 fibers. Appl. Phys. Lett. 2014, 104, 162907. [Google Scholar] [CrossRef]
  2. Cao, J.; Jia, Y.; Wan, X.; Li, B.; Zhang, Y.; Huang, S.; Yang, H.; Yuan, G.; Li, G.; Cui, X.; et al. Strong tribocatalysis of strontium titantate nanofibers through harvesting friction energy for dye decomposition. Ceram. Int. 2022, 48, 9651–9657. [Google Scholar] [CrossRef]
  3. Zheng, Y.; Jia, Y.; Li, H.; Wu, Z.; Dong, X. Enhanced piezo-electro-chemical coupling effect of BaTiO3/g-C3N4 nanocomposite for vibration-catalysis. J. Mater. Sci. 2020, 55, 14787–14797. [Google Scholar] [CrossRef]
  4. Zhao, J.; Chen, L.; Luo, W.; Li, H.; Wu, Z.; Xu, Z.; Zhang, Y.; Zhang, H.; Yuan, G.; Gao, J.; et al. Strong tribo-catalysis of zinc oxide nanorods via triboelectrically-harvesting friction energy. Ceram. Int. 2020, 46, 25293–25298. [Google Scholar] [CrossRef]
  5. Ma, J.; Jia, Y.; Chen, L.; Zheng, Y.; Wu, Z.; Luo, W.; Jiang, M.; Cui, X.; Li, Y. Dye wastewater treatment driven by cyclically heating/ cooling the poled (K0.5Na0.5)NbO3 pyroelectric crystal catalyst. J. Clean. Prod. 2020, 276, 124218. [Google Scholar] [CrossRef]
  6. Chen, L.; Jia, Y.; Zhao, J.; Ma, J.; Wu, Z.; Yuan, G.; Cui, X. Strong piezoelectric-catalysis in barium titanate/carbon hybrid nanocomposites for dye wastewater decomposition. J. Colloid Interf. Sci. 2021, 586, 758–765. [Google Scholar] [CrossRef]
  7. Wang, L.; Haugen, N.O.; Wu, Z.; Shu, X.; Jia, Y.; Ma, J.; Yu, S.; Li, H.; Chai, Q. Ferroelectric BaTiO3@ZnO heterostructure nanofibers with enhanced pyroelectrically-driven-catalysis. Ceram. Int. 2019, 45, 90–95. [Google Scholar] [CrossRef]
  8. Ma, J.; Wu, Z.; Luo, W.; Zheng, Y.; Jia, Y.; Wang, L.; Huang, H. High pyrocatalytic properties of pyroelectric BaTiO3 nanofibers loaded by noble metal under room-temperature thermal cycling. Ceram. Int. 2018, 44, 21835–21841. [Google Scholar] [CrossRef]
  9. Xu, X.; Jia, Y.; Xiao, L.; Wu, Z. Strong vibration-catalysis of ZnO nanorods for dye wastewater decolorization via piezo-electro-chemical coupling. Chemosphere 2018, 193, 1143–1148. [Google Scholar] [CrossRef]
  10. Wu, Z.; Xu, T.; Ruan, L.; Guan, J.; Huang, S.; Dong, X.; Li, H.; Jia, Y. Strong tribo-catalytic nitrogen fixation of two-dimensional graphite carbon nitride through harvesting mechanical energy. Nanomaterials 2022, 12, 1981. [Google Scholar] [CrossRef]
  11. Yao, Y.; Jia, Y.; Zhang, Q.; Li, S.; Li, G.; Cui, X.; Wu, Z. Piezoelectric BaTiO3 with milling treatment for highly efficient piezocatalysis under vibration. J. Alloys Compd. 2022, 905, 164234. [Google Scholar] [CrossRef]
  12. Guan, J.; Jia, Y.; Chang, T.; Ruan, L.; Xu, T.; Zhang, Z.; Yuan, G.; Wu, Z.; Zhu, G. Highly efficient piezo-catalysis of the heat-treated cellulose nanocrystal for dye decomposition driven by ultrasonic vibration. Sep. Purif. Technol. 2022, 286, 120450. [Google Scholar] [CrossRef]
  13. Guan, J.; Jia, Y.; Cao, J.; Yuan, G.; Huang, S.; Cui, X.; Li, G.; Wu, Z. Enhancement of piezoelectric catalysis of Na0.5Bi0.5TiO3 with electric poling for dye decomposition. Ceram. Int. 2022, 48, 3695–3701. [Google Scholar] [CrossRef]
  14. Dong, X.; Cheng, F. Recent development in exfoliated two-dimensional g-C3N4 nanosheets for photocatalytic applications. J. Mater. Chem. A 2015, 3, 23642–23652. [Google Scholar] [CrossRef]
  15. Xu, X.; Xiao, L.; Jia, Y.; Hong, Y.; Ma, J.; Wu, Z. Strong visible light photocatalytic activity of magnetically recyclable sol–gel-synthesized ZnFe2O4 for Rhodamine B degradation. J. Electron. Mater. 2017, 47, 536–541. [Google Scholar] [CrossRef]
  16. Yang, W.; Chen, Y.; Gao, S.; Sang, L.; Tao, R.; Sun, C.; Shang, J.K. Post-illumination activity of Bi2WO6 in the dark from the photocatalytic "memory" effect. J. Adv. Ceramics 2021, 10, 355–367. [Google Scholar] [CrossRef]
  17. Chen, D.; Wang, D.; Ge, Q.; Ping, G.; Fan, M.; Qin, L.; Bai, L.; Lv, C.; Shu, K. Graphene-wrapped ZnO nanospheres as a photocatalyst for high performance photocatalysis. Thin. Solid Films 2015, 574, 1–9. [Google Scholar] [CrossRef]
  18. Fu, J.; Tian, Y.; Chang, B.; Xi, F.; Dong, X. Facile fabrication of N-doped TiO2 nanocatalyst with superior performance under visible light irradiation. J. Solid State Chem. 2013, 199, 280–286. [Google Scholar] [CrossRef]
  19. You, H.; Wu, Z.; Jia, Y.; Xu, X.; Xia, Y.; Han, Z.; Wang, Y. High-efficiency and mechano-/photo-bi-catalysis of piezoelectric-ZnO@photoelectric-TiO2 core-shell nanofibers for dye decomposition. Chemosphere 2017, 183, 528–535. [Google Scholar] [CrossRef]
  20. Xu, X.; Chen, S.; Wu, Z.; Jia, Y.; Xiao, L.; Liu, Y. Strong pyro-electro-chemical coupling of Ba0.7Sr0.3TiO3@Ag pyroelectric nanoparticles for room-temperature pyrocatalysis. Nano Energy 2018, 50, 581–588. [Google Scholar] [CrossRef]
  21. Tang, Y.; Luo, L.; Jia, Y.; Luo, H.; Zhao, X.; Xu, H.; Lin, D.; Sun, J.; Meng, X.; Zhu, J.; et al. Mn-doped 0.71Pb(Mg1/3Nb2/3O3)-0.29PbTiO3 pyroelectric crystals for uncooled infrared focal plane arrays applications. Appl. Phys. Lett. 2006, 89, 162906. [Google Scholar] [CrossRef] [Green Version]
  22. You, H.; Jia, Y.; Wu, Z.; Wang, F.; Huang, H.; Wang, Y. Room-temperature pyro-catalytic hydrogen generation of 2D few-layer black phosphorene under cold-hot alternation. Nat. Commun. 2018, 9, 2889. [Google Scholar] [CrossRef] [Green Version]
  23. Chen, C.; Wang, Y.; Li, J.; Wu, C.; Yang, G. Piezoelectric, ferroelectric and pyroelectric properties of (100-x)Pb(Mg1/3Nb2/3)O3−x PbTiO3 ceramics. J. Adv. Dielectr. 2022, 12, 2250002. [Google Scholar] [CrossRef]
  24. Zhang, B.; Sun, R.; Wang, F.; Feng, T.; Zhang, P.; Luo, H. Pyroelectric properties of 91.5Na0.5Bi0.5TiO3–8.5K0.5Bi0.5TiO3 lead-free single crystal. J. Adv. Dielectr. 2021, 11, 2150023. [Google Scholar] [CrossRef]
  25. Kaur, S.; Arora, M.; Kumar, S.; Malhi, P.S.; Singh, M.; Singh, A. Agreement in experimental and theoretically obtained electrocaloric effect in optimized Bi3+ doped PbZr0.52Ti0.48O3 material. J. Adv. Dielectr. 2022, 12, 2250003. [Google Scholar] [CrossRef]
  26. Najmi, F.; Shen, W.; Cremaschi, L.; Cheng, Z.-Y. Electrocaloric devices part I: Analytical solution of one-dimensional transient heat conduction in a multilayer electrocaloric system. J. Adv. Dielectr. 2020, 10, 2050028. [Google Scholar] [CrossRef]
  27. Najmi, F.; Shen, W.; Cremaschi, L.; Cheng, Z.-Y. Electrocaloric devices part II: All-solid heat pump without moving parts. J. Adv. Dielectr. 2020, 10, 2050029. [Google Scholar] [CrossRef]
  28. Xu, X.; Xiao, L.; Jia, Y.; Wu, Z.; Wang, F.; Wang, Y.; Haugen, N.O.; Huang, H. Pyro-catalytic hydrogen evolution by Ba0.7Sr0.3TiO3 nanoparticles: Harvesting cold–hot alternation energy near room-temperature. Energ. Environ. Sci. 2018, 11, 2198–2207. [Google Scholar] [CrossRef]
  29. Yang, B.; Chen, H.; Yang, Y.; Wang, L.; Bian, J.; Liu, Q.; Lou, X. Insights into the tribo-/pyro-catalysis using Sr-doped BaTiO3 ferroelectric nanocrystals for efficient water remediation. Chem. Eng. J. 2021, 416, 128986. [Google Scholar] [CrossRef]
  30. Min, M.; Liu, Y.; Song, C.; Zhao, D.; Wang, X.; Qiao, Y.; Feng, R.; Hao, W.; Tao, P.; Shang, W.; et al. Photothermally enabled pyro-catalysis of a BaTiO3 nanoparticle composite membrane at the liquid/air Interface. ACS Appl. Mater. Interfaces 2018, 10, 21246–21253. [Google Scholar] [CrossRef]
  31. Liu, Y.; Wang, X.; Qiao, Y.; Min, M.; Wang, L.; Shan, H.; Ma, Y.; Hao, W.; Tao, P.; Shang, W.; et al. Pyroelectric synthesis of metal–BaTiO3 hybrid nanoparticles with enhanced pyrocatalytic performance. ACS Sustain. Chem. Eng. 2019, 7, 2602–2609. [Google Scholar] [CrossRef]
  32. Chen, M.; Jia, Y.; Li, H.; Wu, Z.; Huang, T.; Zhang, H. Enhanced pyrocatalysis of the pyroelectric BiFeO3/g-C3N4 heterostructure for dye decomposition driven by cold-hot temperature alternation. J. Adv. Ceram. 2021, 10, 338–346. [Google Scholar] [CrossRef]
  33. Wu, Z.; Luo, W.; Zhang, H.; Jia, Y. Strong pyro-catalysis of shape-controllable bismuth oxychloride nanomaterial for wastewater remediation. Appl. Surf. Sci. 2020, 513, 145630. [Google Scholar] [CrossRef]
  34. Luo, W.; Ying, J.; Yu, S.; Yang, X.; Jia, Y.; Chen, M.; Zhang, H.; Gao, J.; Li, Y.; Mai, Y.-W.; et al. ZnS:Cu powders with strong visible-light photocatalysis and pyro-catalysis for room-temperature dye decomposition. Ceram. Int. 2020, 46, 12096–12101. [Google Scholar] [CrossRef]
  35. Raufeisen, S.; Neumeister, P.; Buchheim, J.R.; Stelter, M.; Braeutigam, P. Pyrocatalytic oxidation- strong size-dependent poling effect on catalytic activity of pyroelectric BaTiO3 nano- and microparticles. Phys. Chem. Chem. Phys. 2020, 22, 23464–23473. [Google Scholar] [CrossRef] [PubMed]
  36. 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. [Google Scholar] [CrossRef]
  37. Chen, L.; Li, H.; Wu, Z.; Feng, L.; Yu, S.; Zhang, H.; Gao, J.; Mai, Y.-W.; Jia, Y. Enhancement of pyroelectric catalysis of ferroelectric BaTiO3 crystal: The action mechanism of electric poling. Ceram. Int. 2020, 46, 16763–16769. [Google Scholar] [CrossRef]
  38. Qian, W.; Wu, Z.; Jia, Y.; Hong, Y.; Xu, X.; You, H.; Zheng, Y.; Xia, Y. Thermo-electrochemical coupling for room temperature thermocatalysis in pyroelectric ZnO nanorods. Electrochem. Commun. 2017, 81, 124–127. [Google Scholar] [CrossRef]
  39. Xia, Y.; Jia, Y.; Qian, W.; Xu, X.; Wu, Z.; Han, Z.; Hong, Y.; You, H.; Ismail, M.; Bai, G.; et al. Pyroelectrically induced pyro-electro-chemical catalytic activity of BaTiO3 nanofibers under room-temperature cold–hot cycle excitations. Metals 2017, 7, 122. [Google Scholar] [CrossRef] [Green Version]
  40. You, H.; Wu, Z.; Wang, L.; Jia, Y.; Li, S.; Zou, J. Highly efficient pyrocatalysis of pyroelectric NaNbO3 shape-controllable nanoparticles for room-temperature dye decomposition. Chemosphere 2018, 199, 513–537. [Google Scholar] [CrossRef]
  41. Wang, Y.; Dai, X.; Dong, C.; Guo, W.; Xu, Z.; Chen, Y.; Xiang, H.; Zhang, R. Engineering electronic band structure of binary thermoelectric nanocatalysts for augmented pyrocatalytic tumor nanotherapy. Adv. Mater. 2022, 34, 2106773. [Google Scholar] [CrossRef] [PubMed]
  42. Chen, F.; Guo, J.; Meng, D.; Wu, Y.; Sun, R.; Zhao, C. Strong pyro-electro-chemical coupling of elbaite/H2O2 system for pyrocatalysis dye wastewater. Catalysts 2021, 11, 1370. [Google Scholar] [CrossRef]
  43. Min, J.; Hao, H.; Zhao, Y.; Hu, Y.; Huang, Q.; Zhu, F.; Zhang, G.; Bi, J.; Yan, S.; Hou, H. Construction of NaNbO3/CdS nanorods composites with pyroelectric effect for enhanced pyrocatalytic and antibacterial activity under room-temperature cold-hot cycles. Colloid Surf. A 2022, 650, 129529. [Google Scholar] [CrossRef]
  44. Zou, X.; Zhu, R.; Cheng, Z.; Shi, X.; Li, L.; Zhou, Y.; Wang, D.; Wang, W.; Wang, Z.; Shao, Y.; et al. Natural day–night temperature variations driving pyrocatalytic wastewater treatment and CO2 reduction based on room-temperature rhombohedral–tetragonal phase transition of lead-free ferroelectric BZT-0.5BCT fibers. Mater. Chem. Phys. 2022, 291, 126719. [Google Scholar] [CrossRef]
  45. Wang, H.; Jia, Y.; Xu, T.; Shu, X.; He, Y.; Huang, S.; Yuan, G.; Cui, X.; Li, G.; Wu, Z. Barium calcium titanate @carbon hybrid materials for high-efficiency room-temperature pyrocatalysis. Ceram. Int. 2022, 48, 10498–10505. [Google Scholar] [CrossRef]
  46. Raufeisen, S.; Stelter, M.; Braeutigam, P. Pyrocatalysis-The DCF assay as a pH-robust tool to determine the oxidation capability of thermally excited pyroelectric powders. PLoS ONE 2020, 15, e0228644. [Google Scholar] [CrossRef] [PubMed]
  47. 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. [Google Scholar] [CrossRef] [Green Version]
  48. Sharma, M.; Patel, S.; Singh, V.P.; Vaish, R. Pyroelectric energy harvesting for dye decolorization using Ba0.9Ca0.1TiO3 ceramics. J. Appl. Phys. 2020, 128, 095108. [Google Scholar] [CrossRef]
  49. Hu, Y.; Hao, H.; Zhao, Y.; Min, J.; Huang, Q.; Zhong, J.; Zhang, G.; Bi, J.; Yan, S.; Hou, H. Construction of pyroelectrically-driven BiFeO3@CuBi2O4 nanofiber composite catalyst for enhanced pyrocatalytic activities under room-temperature cold and hot cycles. Surf. Interfaces 2022, 33, 102191. [Google Scholar] [CrossRef]
  50. Xiao, L.; Xu, X.; Jia, Y.; Hu, G.; Hu, J.; Yuan, B.; Yu, Y.; Zou, G. Pyroelectric nanoplates for reduction of CO2 to methanol driven by temperature-variation. Nat. Commun. 2021, 12, 318. [Google Scholar] [CrossRef]
  51. Bai, Y.; Zheng, G.-P.; Shi, S.-Q. Abnormal electrocaloric effect of Na0.5Bi0.5TiO3–BaTiO3 lead-free ferroelectric ceramics above room temperature. Mater. Res. Bull. 2011, 46, 1866–1869. [Google Scholar] [CrossRef]
  52. Yan, H.; Xiao, D.; Yu, P.; Zhu, J.; Lin, D.; Li, G. The dependence of the piezoelectric properties on the differences of the A-site and B-site ions for (Bi1−xNax)TiO3-based ceramics. Mater. Design 2005, 26, 474–478. [Google Scholar] [CrossRef]
  53. Smolenskii, G.A.; Isupov, V.A.; Agranovskaya, A.I.; Krainik, N.N. New ferroelectrics of complex composition. Sov. Phys.-Solid State 1961, 2, 2651–2654. [Google Scholar]
  54. Behara, S.; Ghatti, L.; Kanthamani, S.; Dumpala, M.; Thomas, T. Structural, optical, and Raman studies of Gd doped sodium bismuth titanate. Ceram. Int. 2018, 44, 12118–12124. [Google Scholar] [CrossRef]
  55. Khaliq, A.; Sheeraz, M.; Ullah, A.; Seog, H.J.; Ahn, C.W.; Kim, T.H.; Cho, S.; Kim, I.W. Ferroelectric seeds-induced phase evolution and large electrostrain under reduced poling field in bismuth-based composites. Ceram. Int. 2018, 44, 13278–13285. [Google Scholar] [CrossRef]
  56. Cheng, Z.; Zhao, Z. Ink-jet printed BNT thin films with improved ferroelectric properties via annealing in wet air. Ceram. Int. 2018, 44, 10700–10707. [Google Scholar] [CrossRef]
  57. Fleischer, C.; Chatzitakis, A.; Norby, T. Intrinsic photoelectrocatalytic activity in oriented, photonic TiO2 nanotubes. Mat. Sci. Semicon. Proc. 2018, 88, 186–191. [Google Scholar] [CrossRef]
  58. Xu, X.; Xiao, L.; Wu, Z.; Jia, Y.; Ye, X.; Wang, F.; Yuan, B.; Yu, Y.; Huang, H.; Zou, G. Harvesting vibration energy to piezo-catalytically generate hydrogen through Bi2WO6 layered-perovskite. Nano Energy 2020, 78, 105351. [Google Scholar] [CrossRef]
  59. Lei, H.; Wu, M.; Mo, F.; Ji, S.; Dong, X.; Wu, Z.; Gao, J.; Yang, Y.; Jia, Y. Tribo-catalytic degradation of organic pollutants through bismuth oxyiodate triboelectrically harvesting mechanical energy. Nano Energy 2020, 78, 105290. [Google Scholar] [CrossRef]
  60. Li, P.; Wu, J.; Wu, Z.; Jia, Y.; Ma, J.; Chen, W.; Zhang, L.; Yang, J.; Liu, Y. Strong tribocatalytic dye decomposition through utilizing triboelectric energy of barium strontium titanate nanoparticles. Nano Energy 2019, 63, 103832. [Google Scholar] [CrossRef]
  61. Reddy, J.K.; Srinivas, B.; Kumari, V.D.; Subrahmanyam, M. Sm3+-doped Bi2O3 photocatalyst prepared by hydrothermal synthesis. ChemCatChem 2009, 1, 492–496. [Google Scholar]
  62. Lee, Y.; Watanabe, T.; Takata, T.; Kondo, J.N.; Hara, M.; Yoshimura, M.; Domen, K. Preparation and characterization of sodium tantalate thin films by hydrothermal− electrochemical synthesis. Chem. Mater. 2005, 17, 2422–2426. [Google Scholar] [CrossRef]
  63. Sun, Z.H.; Kim, C.H.; Moon, H.B.; Jang, Y.H.; Cho, J.H. Effect of Mn doping on the electronic structures of CaCu3Ti4O12 ceramics. J. Korean Phys. Soc. 2009, 54, 881–885. [Google Scholar] [CrossRef]
  64. 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. [Google Scholar] [CrossRef] [Green Version]
  65. Ma, J.; Chen, L.; Wu, Z.; Chen, J.; Jia, Y.; Hu, Y. Pyroelectric Pb(Zr0.52Ti0.48)O3 polarized ceramic with strong pyro-driven-catalysis for dye wastewater decomposition. Ceram. Int. 2019, 45, 11934–11938. [Google Scholar] [CrossRef]
  66. Wu, J.; Mao, W.; Wu, Z.; Xu, X.; You, H.; Xue, A.X.; Jia, Y. Strong pyro-catalysis of pyroelectric BiFeO3 nanoparticles under a room-temperature cold-hot alternation. Nanoscale 2016, 8, 7343–7350. [Google Scholar] [CrossRef]
  67. Yang, X.Z.; Yang, D.J.; Zhu, H.Y.; Liu, J.W.; Wayde, N.M.; Ray, F.; Lisa, D.; Shen, Y.N. Mesoporous structure with size controllable anatase attached on silicate layers for efficient photocatalysis. J. Phys. Chem. C 2009, 113, 8243–8248. [Google Scholar]
  68. Testino, A.; Bellobono, I.R.; Buscaglia, V.; Canevali, C.; D’Arienzo, M.; Polizzi, S.; Scotti, R.; Morazzoni, F. Optimizing the photocatalytic properties of hydrothermal TiO2 by the control of phase composition and particle morphology. A systematic approach. J. Am. Chem. Soc. 2007, 129, 3564–3575. [Google Scholar] [CrossRef]
  69. Chen, C.; Zhao, W.; Lei, P.; Zhao, J.; Serpone, N. Photosensitized degradation of dyes in polyoxometalate solutions versus TiO2 dispersions under visible-light irradiation: Mechanistic implications. Chem. Eur. J. 2004, 10, 1956–1965. [Google Scholar] [CrossRef]
  70. Wang, C.; Tian, N.; Ma, T.; Zhang, Y.; Huang, H. Pyroelectric catalysis. Nano Energy 2020, 78, 105371. [Google Scholar] [CrossRef]
  71. Zhao, T.; Zai, J.; Xu, M.; Zou, Q.; Su, Y.; Wang, K.; Qian, X. Hierarchical Bi2O2CO3 microspheres with improved visible-light-driven photocatalytic activity. CrystEngComm 2011, 13, 4010. [Google Scholar] [CrossRef]
  72. Li, J.; Wang, G.; Wang, H.; Tang, C.; Wang, Y.; Liang, C.; Cai, W.; Zhang, L. In situ self-assembly synthesis and photocatalytic performance of hierarchical Bi0.5Na0.5TiO3 micro/nanostructures. J. Mater. Chem. 2009, 19, 2253. [Google Scholar] [CrossRef]
  73. Baskar, G.; Aiswarya, R. Trends in catalytic production of biodiesel from various feedstocks. Renew. Sust. Energy Rev. 2016, 57, 496–504. [Google Scholar] [CrossRef]
  74. Jiang, D.; Wang, W.; Gao, E.; Zhang, L.; Sun, S. Bismuth-induced integration of solar energy conversion with synergistic low-temperature catalysis in Ce1–xBixO2−δ nanorods. J. Phys. Chem. C 2013, 117, 24242–24249. [Google Scholar] [CrossRef]
  75. Yu, C.; Tan, M.; Tao, C.; Hou, Y.; Liu, C.; Meng, H.; Su, Y.; Qiao, L.; Bai, Y. Remarkably enhanced piezo-photocatalytic performance in BaTiO3/CuO heterostructures for organic pollutant degradation. J. Adv. Ceram. 2022, 11, 414–426. [Google Scholar] [CrossRef]
  76. He, T.; Cao, Z.; Li, G.; Jia, Y.; Peng, B. High efficiently harvesting visible light and vibration energy in (1−x)AgNbO3–xLiTaO3 solid solution around antiferroelectric–ferroelectric phase boundary for dye degradation. J. Adv. Ceram. 2022, 11, 1641–1653. [Google Scholar] [CrossRef]
  77. Mani, A.D.; Li, J.; Wang, Z.; Zhou, J.; Xiang, H.; Zhao, J.; Deng, L.; Yang, H. Coupling of piezocatalysis and photocatalysis for efficient degradation of methylene blue by Bi0.9Gd0.07La0.03FeO3 nanotubes. J. Adv. Ceram. 2022, 11, 1069–1081. [Google Scholar] [CrossRef]
Figure 1. Microstructure characterization of BNT. (a) The XRD pattern. The inset is the TEM image. (b) EDX analysis. The inset is the SEM image.
Figure 1. Microstructure characterization of BNT. (a) The XRD pattern. The inset is the TEM image. (b) EDX analysis. The inset is the SEM image.
Nanomaterials 12 04091 g001
Figure 2. The XPS spectra of BNT.
Figure 2. The XPS spectra of BNT.
Nanomaterials 12 04091 g002
Figure 3. The XPS spectra of BNT: (a) Bi 4f, (b) Na 1s, (c) Ti 2p and (d) O 1s.
Figure 3. The XPS spectra of BNT: (a) Bi 4f, (b) Na 1s, (c) Ti 2p and (d) O 1s.
Nanomaterials 12 04091 g003
Figure 4. The ferroelectric hysteresis loop of BNT. The inset is the schematic diagram for the ferroelectric measurement.
Figure 4. The ferroelectric hysteresis loop of BNT. The inset is the schematic diagram for the ferroelectric measurement.
Nanomaterials 12 04091 g004
Figure 5. The adsorption spectrum of RhB dye solution with BNT. The inset shows the time curve for the cold-hot cycle.
Figure 5. The adsorption spectrum of RhB dye solution with BNT. The inset shows the time curve for the cold-hot cycle.
Nanomaterials 12 04091 g005
Figure 6. The mechanism diagram of the pyro-catalytic dye decomposition with BNT.
Figure 6. The mechanism diagram of the pyro-catalytic dye decomposition with BNT.
Nanomaterials 12 04091 g006
Figure 7. The pyro-catalytic decomposition ratio of RhB and inhibitory effect of different inhibitors on BNT. The inset shows these pictures of RhB dye decomposition.
Figure 7. The pyro-catalytic decomposition ratio of RhB and inhibitory effect of different inhibitors on BNT. The inset shows these pictures of RhB dye decomposition.
Nanomaterials 12 04091 g007
Figure 8. The characterization of pyro-catalysis performance of BNT: (a) MB and (b) MO dye solutions. The inset shows pictures of MB or MO dye decomposition.
Figure 8. The characterization of pyro-catalysis performance of BNT: (a) MB and (b) MO dye solutions. The inset shows pictures of MB or MO dye decomposition.
Nanomaterials 12 04091 g008
Figure 9. The pyro-catalytic dye decomposition ratios for different dye solutions.
Figure 9. The pyro-catalytic dye decomposition ratios for different dye solutions.
Nanomaterials 12 04091 g009
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wu, Z.; Wu, S.; Hong, S.; Shi, X.; Guo, D.; Zhang, Y.; Xu, X.; Chen, Z.; Jia, Y. Lead-Free Bi0.5Na0.5TiO3 Ferroelectric Nanomaterials for Pyro-Catalytic Dye Pollutant Removal under Cold-Hot Alternation. Nanomaterials 2022, 12, 4091. https://doi.org/10.3390/nano12224091

AMA Style

Wu Z, Wu S, Hong S, Shi X, Guo D, Zhang Y, Xu X, Chen Z, Jia Y. Lead-Free Bi0.5Na0.5TiO3 Ferroelectric Nanomaterials for Pyro-Catalytic Dye Pollutant Removal under Cold-Hot Alternation. Nanomaterials. 2022; 12(22):4091. https://doi.org/10.3390/nano12224091

Chicago/Turabian Style

Wu, Zheng, Siqi Wu, Siqi Hong, Xiaoyu Shi, Di Guo, Yan Zhang, Xiaoli Xu, Zhi Chen, and Yanmin Jia. 2022. "Lead-Free Bi0.5Na0.5TiO3 Ferroelectric Nanomaterials for Pyro-Catalytic Dye Pollutant Removal under Cold-Hot Alternation" Nanomaterials 12, no. 22: 4091. https://doi.org/10.3390/nano12224091

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