Next Article in Journal
Structural Investigations of TiC–Cu Nanocomposites Prepared by Ball Milling and Spark Plasma Sintering
Next Article in Special Issue
A Novel Hybrid Actuator Driven Magnetically in the Bi-Cell PEM Fuel Cell Stack
Previous Article in Journal
Influence of NbC Addition on the Sintering Behaviour of Medium Carbon PM Steels
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Pyroelectrically Induced Pyro-Electro-Chemical Catalytic Activity of BaTiO3 Nanofibers under Room-Temperature Cold–Hot Cycle Excitations

1
Department of Physics, Zhejiang Normal University, Jinhua 321004, China
2
College of Geography and Environmental Sciences, Zhejiang Normal University, Jinhua 321004, China
3
Department of Chemistry, Zhejiang Normal University, Jinhua 321004, China
*
Authors to whom correspondence should be addressed.
Metals 2017, 7(4), 122; https://doi.org/10.3390/met7040122
Submission received: 1 March 2017 / Revised: 24 March 2017 / Accepted: 27 March 2017 / Published: 1 April 2017
(This article belongs to the Special Issue Piezoelectric Materials and Applications)

Abstract

:
A pyro-electro-chemical catalytic dye decomposition using lead-free BaTiO3 nanofibers was realized under room-temperature cold–hot cycle excitation (30–47 °C) with a high Rhodamine B (RhB) decomposition efficiency ~99%, which should be ascribed to the product of pyro-electric effect and electrochemical redox reaction. Furthermore, the existence of intermediate product of hydroxyl radical in pyro-electro-chemical catalytic process was also observed. There is no significant decrease in pyro-electro-chemical catalysis activity after being recycled five times. The pyro-electrically induced pyro-electro-chemical catalysis provides a high-efficient, reusable and environmentally friendly technology to remove organic pollutants from water.

1. Introduction

Organic dyes from printing, textile, and tanning industries are one of the most harmful pollutants in wastewater [1,2]. To decompose those organic dye pollutants, various methods have been developed, such as physical technologies, biological processes and semiconductor-based photo-catalysis. However, there still exist many defects in such technologies. Physical technologies such as adsorption on activated carbon, flocculation and reverse osmosis can only transfer dyes from liquid phase to solid phase apart from the secondary pollution problem [3,4,5]. Most dyes dissolved in water are stable and toxic, which creates difficulties for biological processes [6,7]. Semiconductor-based photo-catalysis using TiO2 was widely studied, whereas it is still limited in practical application due to slow reaction rate and usage limitation [8,9,10]. It is necessary to develop a new environmentally friendly and efficient catalytic process.
Similar to photovoltaic effect, pyro-electric materials can convert temperature oscillation to electrical energy by pyro-electric effect with high thermal-to-electric energy conversion efficiency of 45%–55% [11,12], compared to that of photovoltaic materials (<20%) [13,14]. Therefore, analogy to classical photo-catalysis, a new catalytic process depending on the product of pyro-electric effect and electrochemical redox reaction under cold–hot cycle excitations, was designed on the basis of the principle of product effect [15], which can be called pyro-electro-chemical catalysis.
One of the most popular pyro-electric materials, BaTiO3, was selected to realize pyro-electro-chemical catalytic dye decomposition because of its non-toxic nature and strong pyro-electric response with a pyro-electric coefficient of ~1 × 10−7 C·cm−2 K−1 [16]. In this work, hydrothermally prepared BaTiO3 nanofibers exhibit remarkable pyro-electro-chemical catalytic activity in dye decomposition with ~99% efficiency under room-temperature cold–hot cycle excitations.

2. Materials and Methods

First, 1.88 g titanium dioxide was added into 60 mL 10 M sodium hydroxide solution. Then, the mixture was sealed in 100 mL Teflon-lined stainless steel autoclave heating at 200 °C for 24 h to obtain sodium titanate (Na2Ti3O7). The obtained Na2Ti3O7 powder was soaked in 0.2 M aqueous hydrochloric acid solution for 4 h to obtain hydrogen titanate (H2Ti3O7) nanofibers. The H2Ti3O7 nanofibers were then washed with deionized (DI) water until the PH reached about 7, and subsequently dried at 80 °C for 12 h. The resultant H2Ti3O7 powder was dispersed in 60 mL of 0.05 M Barium hydroxide aqueous solution to fully transfer H2Ti3O7 nanofibers to BaTiO3 nanofibers. The mixture was set in 100 mL Teflon-lined stainless steel to react at 200 °C for 2 h. The obtained precipitate was washed with ethanol and DI water before drying at 60 °C for 12 h.
To investigate the pyro-electro-chemical catalytic activity of BaTiO3 nanofibers, dye decomposition experiments were carried out with rhodamine B (RhB) as model organic pollutant. In the process, 50 mg of as-prepared BaTiO3 nanofibers were added into 50 mL RhB dye solution with initial concentrate of 5 mg/L. Before performing cold–hot cycle excitations, the mixture was stirred for 2 h in darkness to achieve equilibrium adsorption. Then, after finishing the physical absorption in advance, the solution was put into a glass beaker placed in the water bath container for pyro-electro-chemical catalysis experiment. The temperature variation (30–47 °C) was applied via water bath container with a slow magnetic stirring (RCT-B-S25, IKA, Königswinter, Germany) in a dark environment to avoid photo-degradation. At regular intervals, 3 mL of solution sample was collected and centrifuged to determine RhB concentration using a U-3900 UV-Vis spectrophotometer (Hitachi, Tokyo, Japan). For comparison purposes, decomposition experiment without BaTiO3 nanofibers was thus carried out.
Radical trapping experiment was performed to explore the mechanism of pyro-electro-chemical catalysis. The hydroxyl (•OH) radicals were detected via fluorescence method employing terephthalic acid for selectively reaction with the generated •OH radicals to produce 2-hydroxyterephthalic acid, which can be detected via a fluorescence spectrometer [17]. Fifty mg of BaTiO3 nanofibers was added into 50 mL, 1 mM of terephthalic acid and NaOH solution to perform the same cold–hot cycle excitations. At regular intervals, 3 mL of mixture was collected and centrifuged to determine the concentration of •OH radicals.
To determine the reusability of the as-prepared BaTiO3 nanofibers, dye decomposition experiment was repeated five times. After performing 72 cold–hot cycle excitations, the mixture was centrifuged to separate out BaTiO3 nanofibers. Then, the obtained nanopowder was washed with DI water and dried at 70 °C for 4 h for the next recycling utilization.
The structure of BaTiO3 nanofibers was investigated using an X-ray diffractometer (PW3040/60, Philips, Eindhoven, the Netherlands). Scanning electron microscopy ( S-4800, Hitachi, Tokyo, Japan) was applied to characterize the surface morphology of the as-prepared nanopowder. Pyro-electro-chemical dye decomposition results were investigated through absorption spectra using a UV-Vis spectrophotometer (U-3900, Hitachi, Tokyo, Japan). Fluorescence spectra were obtained via a fluorescence spectrometer (F-4500, Hitachi, Tokyo, Japan).

3. Results

The X-ray diffraction (XRD) pattern of the as-prepared nanopowder is shown in Figure 1. The XRD pattern displays strong peak intensity, which means high crystallites of the as-prepared powder. It can be seen that this XRD pattern can be indexed to the pure tetragonal phase of BaTiO3 (JCPDS, 05-0526) having no other impurity peaks.
Figure 2 is the scanning electron microscope (SEM) result of BaTiO3 nanofibers, which has an average diameter of 200 nm and an aspect ratio of ~50.
Our pyro-electro-chemical catalysis is originated from the ferro-/pyroelectric effect. In general, the ferro-/piezoelectric performance of the aligned nanofiber ferroelectric arrays can be characterized through Piezoresponse Force Microscopy (PFM) [18,19,20]. However, unlike arrays, our as-prepared BaTiO3 nanofiber powders are orderless. It is difficult to directly measure ferro-/piezoelectric performance. During the growth, the sub-crystal grows in a certain crystallographic direction and the resultant as-prepared pyroelectric nanofibers are in single-crystalline form [21,22,23], which means that the electric dipoles that exist in pyroelectric nanofibers are monotonous. Therefore, hydrothermally prepared pyroelectric nanofibers behave as ferro-/pyroelectric effect without undergoing poling process [24]. It has been reported that BaTiO3 nanofiber has a high piezoelectric coefficient (d33) of ~45 pC/N [16,25].
Figure 3 shows the temperature curve of one cold–hot cycle excitation. Every cold–hot cycle includes 5 min heating time, 5 min cooling time and 5 min hot/cold temperature keeping time.
In pyro-electro-catalysis, the pyroelectrically induced electric potential difference between the anode and cathode is critical for electrochemical catalytic process and the necessary potential difference is thermodynamically at least 1.7 V and 1.9 V for the generation of •OH and •O2 radicals, respectively [26,27]. Both the temperature alternation range and the size of pyroelectric particle can affect the output electric potential. In general, the pyroelectric BaTiO3 nanofibers with these sizes of micrometer in length can generate 0.4 V for 1 °C temperature change [28]. In our experiment, cold–hot cycle excitation range varies from 30 to 47 °C. Therefore, the estimated electric potential (~6.8 V) is enough to drive our pyro-electro-chemical catalysis. Generally, the size of nano-catalyst materials will obviously influence the catalytic oxidation degradation efficiency [29]. Decreasing the sizes of nano-material is helpful for electric charges’ quick migration between catalyst and pollutants, resulting in a high degradation efficiency. However, taking the lowest electric potential demand for generation of some strong oxidation actives into account, it is better to adopt pyroelectric nanofibers, not nanoparticles, to design pyro-electro-chemical catalysis.
The adsorption spectra of RhB are shown in Figure 4. The adsorption peak intensity decreases continuously with the increasing cold–hot cycle times, which results in the decreases of RhB concentration. After performing 72 cold–hot cycle excitations, the adsorption peak disappears.
The principle of pyro-electro-chemical catalysis can be described as in Figure 5: With temperature gradient (ΔT), pyro-electric materials have the ability to achieve thermal-to-electric conversion. In pyro-electric material, the spontaneous polarization (Ps) dipole moment is non-zero, which results in the existence of bound polarization charges in the two opposite sides of the inner surface. When the temperature comes to a constant, the bound polarization charges are shielded by these equal and opposite absorbed charges on the surface of pyro-electric material; the result is electrically neutral [30]. When pyro-electric material is heated, the Ps density decreases as dipole moments begin to lose their orientation. Charge compensation occurs to balance the extra absorbed charges on the surface caused by the decrease of the Ps density [31], which results in the formation of •O2 radicals. When pyro-electric material is cooled, the negative temperature gradient leads to an increase in Ps density as dipole moments begin to regain their orientation, which causes the charges adsorption and the generation of •OH radicals on the surface. Chain reactions of those strong oxidative •O2 and •OH radicals occur to decompose the dye solution [32]. Based on the mechanism analysis, the chemical equations describing the pyro-electro-chemical dye decomposition can be drawn as:
BaTi O 3 + Δ T pyroelectricity q + + q
OH + q + OH
O 2 + q O 2
Dye + OH   ( or O 2 ) Decomposition
Figure 6 shows the RhB decomposition results with and without BaTiO3 nanofibers under cold–hot cycle excitations. The decomposition efficiency (D) can be calculated as:
D = C 0 C t C 0 × 100 %
where C0 and Ct are the initial concentration and actual concentration of RhB after performing different cold–hot cycle excitations, respectively. It can be observed that BaTiO3 nanofibers possess a remarkable pyro-electro-chemical catalytic effect in RhB decomposition with a high efficiency of ~99% after performing 72 cold–hot cycle excitations. The inset in Figure 6 shows the photograph of RhB solutions after different cold–hot cycle times. The color of RhB fades with the increasing cold–hot cycle times and finally disappeared after 72 cold–hot cycle times, which indicates the accomplishment of RhB decomposition. However, the RhB concentration changed slightly without BaTiO3 nanofibers under cold–hot cycle excitations. These dye decomposition experimental results demonstrate that the pyro-electro-chemical catalysis should be ascribed to the pyro-electric effect of BaTiO3 nanofibers, not the pyrolysis of organic dye occurring at high temperatures of 400–1000 °C [33]. Besides pyro-electric materials, cold–hot cycle excitations are also necessary for the achievement of pyro-electro-chemical catalysis. In general, the pyro-electric coefficient (α = ∂Ps/∂T) may be described as a change in the Ps with temperature. The pyro-electric effect is thus related to temperature variation, which means that without cold–hot cycle excitations, it is difficult to realize pyro-electro-chemical catalysis of pyro-electric materials.
To further investigate the role of pyro-electro-chemical catalysis, radical trapping experiment was carried out using terephthalic acid as •OH radical agent. Figure 7 shows fluorescence spectra of terephthalic acid solution after performing different times of cold–hot cycle under an excitation wavelength of 315 nm. Remarkable characteristic fluorescence emission peak signals of 2-hydroxyterephthalic acid are observed after performing cold–hot cycle excitations, which is a result of the selective reaction between terephthalic acid and •OH radicals. The intensity of characteristic peaks at ~425 nm increases almost linearly with the thermal cycles as shown in the inset of Figure 7, suggesting •OH stable production during the pyro-electro-catalytic RhB dye decomposition.
In general, thermal cycle times can affect the decomposition efficiency D. With the increase of thermal cycle times, the generation of electric charges also increases, resulting in the increase of the number of strong oxidizing hydroxyl and the increase of D as shown in Figure 7. The volume of wastewater or the content of catalyst can also affect the decomposition efficiency D. Our pyro-electro-chemical catalysis is the product of pyroelectric effect and electrochemical redox reaction. Being similar to photo-catalysis, volume of wastewater mainly affects the electrochemical redox reaction of the pyro-electro-chemical catalysis process. In theory, the dependence of volume of wastewater on dye decomposition efficiency in pyro-electro-chemical catalysis should have a similar trend with that in photo-catalysis. In general, D will decrease for the excess volume of wastewater, which has been widely reported [34,35].
To investigate the reusability and stability of pyro-electro-chemical catalysis in our as-prepared BaTiO3 pyroelectric fibers, a recycling utilization experiment testing pyro-electro-catalytic RhB decomposition ability of the as-prepared BaTiO3 nanofibers was performed as shown in Figure 8. It can be observed that BaTiO3 nanofibers exhibit little reduced pyro-electro-catalytic activity in dye decomposition with a reduced efficiency of ~3% after being recycled for five dye decomposition experiments, which means that BaTiO3 nanofibers can be repeatedly heated and cooled without losing too much catalytic activity. It reveals the outstanding recyclability and stability of BaTiO3 nanofibers, which is beneficial for its long-term use in decomposition of organic pollutants.

4. Conclusions

In this work, pyro-electro-chemical dye decomposition using hydrothermally prepared BaTiO3 nanofibers has been achieved. The decomposition is high, up to ~99% within 72 cold–hot cycle excitations. The intermediate product of the pyro-electro-chemical catalysis was detected by radical trapping experiment. There is no significant decrease in pyro-electro-chemical catalysis activity after being recycled five times. The pyro-electro-chemical catalysis using lead-free BaTiO3 nanofibers offers a high-efficient, reusable and environmentally friendly technology to decompose organic dye.

Acknowledgments

The work was supported by the National Natural Science Foundation of China (51502266); Weiqi Qian thanks the support of the National College Students’ Innovation Undertaking Training Project of China (No. 201610345021); Zichen Han thanks the support of the Undergraduate’s Extra-Curricular Academic Activities Key Project of College of Geography and Environmental Sciences of Zhejiang Normal University (No. 9).

Author Contributions

Yanmin Jia and Zheng Wu conceived and designed the experiments; Yuntao Xia and Xiaoli Xu performed the pyro-electro-chemical catalytic experiments and recycling performance; Zichen Han and Huilin You measured the XRD and SEM; Ge Bai and Liwei Wang measured the strong oxidation actives; Muhammad Ismail described the physical mechanism of pyro-electro-chemical catalysis of pyroelectric nanofibers. Weiqi Qian and Yuanting Hong analyzed the data; Yuntao Xia wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest. The founding sponsors had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; and in the decision to publish the results.

References

  1. Gupta, K.V.; Pathania, D.; Agarwal, S.; Singh, P. Adsorptional photocatalytic degradation of methylene blue onto pectin–CuS nanocomposite under solar light. J. Hazard. Mater. 2012, 243, 179–186. [Google Scholar] [CrossRef] [PubMed]
  2. Lachheb, H.; Puzenat, E.; Houas, A.; Ksibi, M.; Elaoui, E.; Guillard, C.; Herrmann, J.M. Photocatalytic degradation of various types of dyes (Alizarin S, Crocein Orange G, Methyl Red, Congo Red, Methylene Blue) in water by UV-irradiated titania. Appl. Catal. B 2002, 39, 75–90. [Google Scholar] [CrossRef]
  3. Zhang, J.; Wu, Z.; Jia, Y.M.; Kan, J.W.; Cheng, G.M. Piezoelectric bimorph cantilever for vibration-producing-hydrogen. Sensors 2013, 13, 367–374. [Google Scholar] [CrossRef] [PubMed]
  4. Moghaddam, S.S.; Moghaddam, M.R.A.; Arami, M. Coagulation/flocculation process for dye removal using sludge from water treatment plant: Optimization through response surface methodology. J. Hazard. Mater. 2010, 175, 651–657. [Google Scholar] [CrossRef] [PubMed]
  5. Liu, M.; Lu, Z.; Chen, Z.; Yu, S.; Gao, C. Comparison of reverse osmosis and nanofiltration membranes in the treatment of biologically treated textile effluent for water reuse. Desalination 2011, 281, 372–378. [Google Scholar] [CrossRef]
  6. Patil, S.S.; Shinde, V.M. Biodegradation studies of aniline and nitrobenzene in aniline plant waste water by gas chromatography. Environ. Sci. Technol. 1988, 22, 1160–1165. [Google Scholar] [CrossRef] [PubMed]
  7. Moore, A.T.; Vira, A.; Fogel, S. Biodegradation of trans-1,2-dichloroethylene by methane-utilizing bacteria in an aquifer simulator. Environ. Sci. Technol. 1989, 23, 403–406. [Google Scholar] [CrossRef]
  8. Hadjiivanov, K.I.; Klissurski, D.K. Surface chemistry of titania (anatase) and titania-supported catalysts. Chem. Soc. Rev. 1996, 25, 61–69. [Google Scholar] [CrossRef]
  9. Heller, A. Chemistry and applications of photocatalytic oxidation of thin organic films. Acc. Chem. Res. 1995, 28, 503–508. [Google Scholar] [CrossRef]
  10. Linsebigler, A.L.; Lu, G.; Yates, J.T. Photocatalysis on TiO2 surfaces: Principles, mechanisms, and selected results. Chem. Rev. 1995, 95, 735–758. [Google Scholar] [CrossRef]
  11. Sebald, G.; Guyomar, D.; Agbossou, A. On thermoelectric and pyroelectric energy harvesting. Smart Mater. Struct. 2009, 18, 125006. [Google Scholar] [CrossRef]
  12. Tang, Y.X.; Luo, L.L.; Jia, Y.M.; Luo, H.S.; Zhao, X.Y.; Xu, H.Q.; Lin, D.; Sun, J.L.; Meng, X.J.; Zhu, J.H.; et al. Mn-doped 0.71Pb(Mg1/3Nb2/3O3)–0.29PbTiO3pyroelectric crystals for uncooled infrared focal plane arrays applications. Appl. Phys. Lett. 2006, 89, 162906. [Google Scholar] [CrossRef]
  13. Nazeeruddin, M.K.; Pechy, P.; Renouard, T.; Zakeeruddin, S.M.; Robin, H.-B.; Comte, P.; Liska, P.; Cevey, L.; Costa, E.; Shklover, V.; et al. Engineering of efficient panchromatic sensitizers for nanocrystalline TiO2-based solar cells. J. Am. Chem. Soc. 2001, 123, 1613–1624. [Google Scholar] [CrossRef] [PubMed]
  14. Grätzel, M. Photoelectrochemical cells. Nature 2001, 414, 338–344. [Google Scholar] [CrossRef] [PubMed]
  15. Jia, Y.M.; Luo, H.; Zhao, X.; Wang, F. Giant magnetoelectric response from a piezoelectric/magnetostrictive laminated composite combined with a piezoelectric transformer. Adv. Mater. 2008, 20, 4776–4779. [Google Scholar] [CrossRef]
  16. Zhang, J.; Wang, C.; Bowen, C. Piezoelectric effects and electromechanical theories at the nanoscale. Nanoscale 2014, 6, 13314–13327. [Google Scholar] [CrossRef] [PubMed]
  17. Ishibashi, K.; Fujishima, A.; Watanabe, T.; Hashimoto, K. Detection of active oxidative species in TiO2 photocatalysis using the fluorescence technique. Electrochem. Commun. 2000, 2, 207–210. [Google Scholar] [CrossRef]
  18. Zhou, Z.; Tang, H.; Sodano, H.A. Vertically aligned arrays of BaTiO3 nanowires. Appl. Mater. Inter. 2013, 5, 11894–11899. [Google Scholar] [CrossRef] [PubMed]
  19. Im, B.; Jun, H.; Lee, K.H.; Lee, S.H.; Yang, I.K.; Jeong, Y.H.; Lee, J.S. Fabrication of a vertically aligned ferroelectric perovskite nanowire array on conducting substrate. Chem. Mater. 2010, 22, 4806–4813. [Google Scholar] [CrossRef]
  20. Zhang, X.Y.; Zhao, X.; Lai, C.W.; Wang, J.; Tang, X.G.; Dai, J.Y. Synthesis and piezoresponse of highly ordered Pb (Zr0.53Ti0.47) O3 nanowire arrays. Appl. Phys. Lett. 2004, 85, 4190–4192. [Google Scholar] [CrossRef]
  21. Wang, Z.; Hu, J.; Yu, M.F. One-dimensional ferroelectric monodomain formation in single crystalline BaTiO3 nanowire. Appl. Phys. Lett. 2006, 89, 263119. [Google Scholar] [CrossRef]
  22. Koka, A.; Sodano, H.A. High-sensitivity accelerometer composed of ultra-long vertically aligned barium titanate nanowire arrays. Nat. Commun. 2013, 4, 2682. [Google Scholar] [CrossRef] [PubMed]
  23. Xu, S.; Hansen, B.J.; Wang, Z.L. Piezoelectric-nanowire-enabled power source for driving wireless microelectronics. Nat. Commun. 2010, 1, 93. [Google Scholar] [CrossRef] [PubMed]
  24. Yu, D.; Zhao, M.; Wang, C.; Wang, L.; Su, W.; Gai, Z.; Wang, C.; Li, J.; Zhang, J. Enhanced piezoelectricity in plastically deformed nearly amorphous Bi12TiO20-BaTiO3 nanocomposites. Appl. Phys. Lett. 2016, 109, 032904. [Google Scholar] [CrossRef]
  25. Wang, F.; Mai, Y.W.; Wang, D.; Ding, R.; Shi, W. High quality barium titanate nanofibers for flexible piezoelectric device applications. Sens. Actuators A Phys. 2015, 233, 195–201. [Google Scholar] [CrossRef]
  26. Zhang, F.; Zhang, C.L.; Wang, W.N.; Cong, H.P.; Qian, H.S. Titanium dioxide/upconversion nanoparticles/cadmium sulfide nanofibers enable enhanced full-spectrum absorption for superior solar light driven photocatalysis. ChemSusChem 2016, 9, 1449–1454. [Google Scholar] [CrossRef] [PubMed]
  27. Zhang, F.; Zhang, C.L.; Peng, H.Y.; Cong, H.P.; Qian, H.S. Near-infrared photocatalytic upconversion nanoparticles/TiO2 nanofibers assembled in large scale by electrospinning. Part. Part. Syst. Charact. 2016, 33, 248–253. [Google Scholar] [CrossRef]
  28. Xie, M.; Dunn, S.; Boulbar, E.L.; Bowen, C.R. Pyroelectric energy harvesting for water splitting. Int. J. Hydrog. Energy 2017. [Google Scholar] [CrossRef]
  29. Li, Y.F.; Liu, Z.P. Particle size, shape and activity for photocatalysis on titania anatase nanoparticles in aqueous surroundings. J. Am. Chem. Soc. 2011, 133, 15743–15752. [Google Scholar] [CrossRef] [PubMed]
  30. Lin, H.; Wu, Z.; Jia, Y.M.; Li, W.J.; 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]
  31. Kalinin, S.V.; Bonnel, D.A. Screening phenomena on oxide surfaces and its implications for local electrostatic and transport measurements. Nano Lett. 2004, 4, 555–560. [Google Scholar] [CrossRef]
  32. Wu, J.M.; Chen, Y.R. Ultraviolet-light-assisted formation of ZnO nanowires in ambient air: Comparison of photoresponsive and photocatalytic activities in zinc hydroxide. J. Phys. Chem. C 2011, 115, 2235–2243. [Google Scholar] [CrossRef]
  33. Konieczny, A.; Mondal, K.; Wiltowski, T.; Dydo, P. Catalyst development for thermocatalytic decomposition of methane to hydrogen. J. Hydrog. Energy 2008, 33, 264–272. [Google Scholar] [CrossRef]
  34. Wu, J.; Mao, W.; Wu, Z.; Xu, X.; You, H.; Jia, Y. Strong pyro-catalysis of pyroelectric BiFeO3 nanoparticles under a room-temperature cold–hot alternation. Nanoscale 2016, 8, 7343–7350. [Google Scholar] [CrossRef] [PubMed]
  35. Chen, P.; Zhang, Y.; Zhao, F.; Gao, H.; Chen, X. Facile microwave synthesis and photocatalytic activity of monodispersed BaTiO3 nanocuboids. Mater. Charact. 2016, 114, 243–253. [Google Scholar] [CrossRef]
Figure 1. The X-ray diffraction pattern of BaTiO3 nanofibers.
Figure 1. The X-ray diffraction pattern of BaTiO3 nanofibers.
Metals 07 00122 g001
Figure 2. The scanning electron microscope image of BaTiO3 nanofibers.
Figure 2. The scanning electron microscope image of BaTiO3 nanofibers.
Metals 07 00122 g002
Figure 3. The temperature curve of a single cold–hot cycle excitation.
Figure 3. The temperature curve of a single cold–hot cycle excitation.
Metals 07 00122 g003
Figure 4. The adsorption spectra of RhB after performing different numbers of cold–hot cycle excitations.
Figure 4. The adsorption spectra of RhB after performing different numbers of cold–hot cycle excitations.
Metals 07 00122 g004
Figure 5. The principle scheme of pyro-electro-chemical catalysis of pyroelectric materials under thermal cycle excitations.
Figure 5. The principle scheme of pyro-electro-chemical catalysis of pyroelectric materials under thermal cycle excitations.
Metals 07 00122 g005
Figure 6. The decomposition efficiency of RhB with or without BaTiO3 nanofibers. The inset is the degradation photograph.
Figure 6. The decomposition efficiency of RhB with or without BaTiO3 nanofibers. The inset is the degradation photograph.
Metals 07 00122 g006
Figure 7. Fluorescence spectra for •OH trapping. The inset is fluorescence intensity peaked at 425 nm against the thermal cycle times.
Figure 7. Fluorescence spectra for •OH trapping. The inset is fluorescence intensity peaked at 425 nm against the thermal cycle times.
Metals 07 00122 g007
Figure 8. Recycling utilization of pyro-electro-catalytic BaTiO3 nanofibers.
Figure 8. Recycling utilization of pyro-electro-catalytic BaTiO3 nanofibers.
Metals 07 00122 g008

Share and Cite

MDPI and ACS Style

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. https://doi.org/10.3390/met7040122

AMA Style

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(4):122. https://doi.org/10.3390/met7040122

Chicago/Turabian Style

Xia, Yuntao, Yanmin Jia, Weiqi Qian, Xiaoli Xu, Zheng Wu, Zichen Han, Yuanting Hong, Huilin You, Muhammad Ismail, Ge Bai, and et al. 2017. "Pyroelectrically Induced Pyro-Electro-Chemical Catalytic Activity of BaTiO3 Nanofibers under Room-Temperature Cold–Hot Cycle Excitations" Metals 7, no. 4: 122. https://doi.org/10.3390/met7040122

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