Next Article in Journal
Suppression of Secondary Electron Emissions on the Graphene-Coated Polyimide Materials Prepared by Chemical Vapor Deposition
Next Article in Special Issue
Component Engineering of Multiphase Nickel Sulfide-Based Bifunctional Electrocatalysts for Efficient Overall Water Splitting
Previous Article in Journal
Study on the Rapid Degradation Performance of Salix/Wheat Straw Fiber Degradable Film
Previous Article in Special Issue
Drive Type Virtual Reality Image on a Head-Mounted Display
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tribo-Catalytic Degradation of Methyl Orange Solutions Enhanced by Silicon Single Crystals

1
Key Laboratory of Artificial Micro- and Nano-Structures of Ministry of Education, School of Physics and Technology, Wuhan University, Wuhan 430072, China
2
Department of Chemistry, University College London, London WC1H 0AJ, UK
3
College of Physics, Chongqing University, Chongqing 401331, China
4
Center of Quantum Materials and Devices, Chongqing University, Chongqing 401331, China
5
Hubei Key Laboratory of Radiation Chemistry and Functional Materials, School of Nuclear Technology and Chemistry and Biology, Hubei University of Science and Technology, Xianning 437100, China
6
Guangdong Provincial Key Laboratory of Metal Toughening Technology and Application, Institute of New Materials, Guangdong Academy of Sciences, Guangzhou 510650, China
*
Author to whom correspondence should be addressed.
Coatings 2023, 13(10), 1804; https://doi.org/10.3390/coatings13101804
Submission received: 21 September 2023 / Revised: 17 October 2023 / Accepted: 18 October 2023 / Published: 20 October 2023
(This article belongs to the Special Issue Advanced Materials for Electrocatalysis and Energy Storage)

Abstract

:
Coating materials on the bottoms of reactors/beakers has emerged as an effective method to regulate tribo-catalytic reactions. In this study, silicon single crystals were coated on the bottoms of glass beakers, in which 30 mg/L methyl orange (MO) solutions suspended with alumina nanoparticles were subjected to magnetic stirring using Teflon magnetic rotary disks. With a gentle rotating speed of 400 rpm for the Teflon disks, the MO solutions were changed from yellow to colorless and the characteristic absorption peak of MO at 450 nm in the UV-Vis spectrum disappeared entirely within 120 min. Mass spectrometry tests were further performed to gain insights into the degradation process, which suggested that the degradation was initiated with the cleavage of the nitrogen-nitrogen double bond in ionized MO molecules by the attack of •OH radicals. Through comparison experiments, we established that the observed degradation was related to the friction between alumina and silicon during magnetic stirring, and hydroxyl and superoxide radicals were formed from the friction, according to electron paramagnetic resonance analysis. It is proposed that electron-hole pairs are excited in silicon single crystals through friction with alumina, which diffuse to the surface of the single crystals and result in the degradation.

1. Introduction

The emergence of tribo-catalysis as a promising approach to address the crises of fossil energy shortage and environmental pollution has captured the attention of many researchers [1,2,3,4,5,6,7]. Most current related research works utilize a standard or modified magnetic stirring system, where frictions between the magnetic rotation section, catalyst, and the bottom of the container play a critical role in converting mechanical energy to chemical energy [8,9,10,11,12,13]. Mechanical energy has been harnessed through friction for applications such as hydrogen generation, dye degradation [14,15,16,17], toxins degradation [18], and CO2 reduction [19,20].
To boost the efficiency of tribo-catalysis in various applications, numerous studies have been conducted in recent years, which can be broadly categorized into two groups. In the first group, investigations are focused on some specific aspects of catalysts in tribo-catalysis, including the synthesis of particles with particular morphologies [9,11,21,22], the modulation of electronic structure [8,18,23,24], and the construction of heterojunctions [10,25]. The other group pays considerable attention to regulating friction pairs in tribo-catalysis. Rao et al. found that after 8 h of magnetic stirring, the degradation efficiencies of Rhodamine B (RhB) solution by BaTiO3 nanoparticles were 88%, 94.7%, and 99.1% for the friction pairs of glass beaker-PTFE bar, PP beaker-PTFE bar, and PTFE beaker-PTFE bar, respectively [23]. Dong et al. discovered that the degradation efficiency of RHB or 2,4-DCP by BiWO3 is significantly improved by adding PP or PTFE particles to form new friction pairs with BiWO3 [26]. In our recent paper on the tribo-catalytic conversion of H2O and CO2 by NiO particles, we found that the production of CH4 increased to 7 and 5 folds when PVC and stainless steel 316 were coated on the reactor bottom, respectively [27]. In another latest paper on the tribo-catalytic conversion of H2O and CO2 by Co3O4 nanoparticles, the amounts of H2 and CH4 increased by 2 and 26 folds, respectively, through coating Ti on the glass reactor bottom [28]. Similarly, for tribo-catalytic conversion of H2O and CO2 using a copper magnetic rotary disk, the production of flammable gases also obviously changed while coating Al2O3, copper, or titanium on reactor bottoms [29]. Regulating friction pair is especially effective in boosting tribo-catalysis efficiency, and coating certain materials on container bottoms is a straightforward method to realize it.
To date, two distinct mechanisms have been proposed for tribo-catalysis—electron transfer across atoms and electron transition [30]. In the former, electrons are transferred between materials through friction, and materials that gain or lose electrons generate active species to initiate subsequent redox reactions. In the latter, electron-hole pairs are excited in a material by mechanical energy absorbed through friction, which then results in redox reactions in the surrounding environments, similar to what happens in photo-catalysis. It is well-known that silicon is the most abundant element in Earth’s crust. It is cost-effective, possesses a narrow band gap, and exhibits excellent processing performance, making it widely adopted in photovoltaic applications to convert solar energy into electricity. According to the second mechanism for tribo-catalysis, the generation of electrons and holes in a material by mechanical energy is primarily determined by its energy band structure. The purpose of this work is to explore tribo-catalysis by materials with narrow band gaps. Given that silicon is a typical semiconductor with a narrow band gap, we coated silicon single crystals on the bottoms of glass beakers, where some organic dyes were degraded by oxide particles through magnetic stirring. With such silicon single crystal coatings, 30 mg/L methyl orange (MO) solutions were found to be quickly de-colorized, even by alumina particles, under magnetic stirring. MO is famous for the presence of high-energy bonds (C=N, N=N) in its molecules, and such a concentrated MO solution is difficult to degrade through a conventional catalytic method [31,32,33].

2. Materials and Methods

2.1. Materials Information

Silicon (110) single crystal wafers (ρ > 105 Ω·cm) with a diameter of 40 mm and a thickness of 0.5 mm were purchased from Hangzhou Jingxin Electronic Technology Co., Hangzhou, China. High-purity α-Al2O3 nanoparticles (99.9 wt%, average particle size: 150–500 nm) were obtained from XFNANO Materials Tech. Co., Ltd., Nanjing, China, and α-Al2O3 laminated powder was purchased from Naiou Nano Technology Co., Shanghai, China.

2.2. Coating Silicon Single Crystal Wafers on the Bottoms of Glass Beakers

For some commercial flat-bottomed glass beakers of ϕ 45 mm × 60 mm, silicon (110) single crystal wafers of ϕ 40 mm × 0.5 mm were first coated on their bottoms through a glue. In this way, flat-bottomed glass beakers with both glass and Si-coated bottoms were available separately for further investigations.

2.3. Tribo-Catalytic Degradation of MO Solutions in Glass Beakers

In a typical experiment, 300 mg of α-Al2O3 was dispersed in 30 mL of 30 mg/L MO aqueous solution in a glass beaker. The suspension was magnetically stirred using a homemade PTFE magnetic rotary disk at 400 rpm in the dark at room temperature (25 °C). The details of the PTFE magnetic rotary disk were described in previous work [15]. During the tribo-catalytic process, 1 mL of the solution was taken out every 30 min, followed by centrifugal separation to obtain the supernatant. The concentration of MO was measured by recording the absorption spectra using a Shimadzu 2550 UV-Vis spectrometer (UV-2550; Shimadzu, Kyoto, Japan) over a 200–800 nm range.

2.4. Analyses of Degradation Products of MO through a Mass Spectrometer

The products resulting from the tribo-catalytic degradation of MO solutions in Si-coated beakers were further analyzed using a mass spectrometer Thermo Q-Exactive Plus (Thermo Scientific, San Jose, CA, USA) equipped with a heated electrospray ionization (HESI) source with the mass scan range set to m/z 60–350. The mobile phases used in the analysis were acetonitrile and 0.01 mol/L acetic acid.

2.5. Detection of Radical Species

Electron paramagnetic resonance (EPR) spectroscopy was employed to probe the reactive oxygen species of •OH and •O2, which are essential in attacking dye macromolecules during the catalytic process of dye degradation. In a Si-coated flat-bottomed beaker, 0.15 g of Al2O3 nanoparticles and 50 μL of 5,5-Dimethyl-1-pyrroline N-oxide (DMPO) were added to 10 mL of deionized water for the detection of hydroxyl radical production; 0.15 g of Al2O3 nanoparticles and 50 μL of 5,5-Dimethyl-1-pyrroline N-oxide (DMPO) were added to 10 mL of methanol for the detection of superoxide radical production. The same experimental conditions were followed to detect •OH and •O2 without Al2O3 nanoparticles. Magnetic stirring was carried out using a PTFE magnetic rotary disk at 400 rpm for 30 min at room temperature without light. EPR spectra were recorded on a Bruker A300 paramagnetic resonance spectrometer.

3. Results and Discussion

The influence of silicon single crystals on tribo-catalytic degradation of various organic dyes employing different catalysts has been thoroughly explored. The tribo-catalytic degradation of 30 mg/L MO aqueous solutions by Al2O3 powders in Si-coated glass beakers has drawn our attention for two compelling reasons. Firstly, Al2O3 is well recognized for its wide band gap, and the excitation of electron-hole pairs through mechanical energy can be attributed to silicon only when it interacts with Al2O3 through friction. Secondly, owing to the presence of azo double bonds in its molecular structure, MO, especially at a high concentration of 30 mg/L, poses a challenging target for degradation. This demonstrates the potential of silicon in tribo-catalysis.
In this study, two types of α-Al2O3 white powders were analyzed using a scanning electron microscope (Zeiss GeminiSEM 500, Jena, Germany), and their SEM images are presented in Figure 1. For the α-Al2O3 nanoparticles, referred to as Al2O3 I hereafter, the particles appear as irregular polyhedrons ranging from 200 to 500 nm, as depicted in Figure 1a. In contrast, the α-Al2O3 laminated powder (Al2O3 II) consists of laminate-like particles as large as 5 μm, as shown in Figure 1b. The two kinds of α-Al2O3 white powders differ greatly in both size and shape.
For reference, Figure 2a displays the UV-visible absorption spectrums of a 30 mg/L MO solution during magnetic stirring with Al2O3 I activated by a Teflon magnetic rotary disk in a glass beaker. No discernible change was observed in the absorption spectrum, even after 150 min of magnetic stirring. This observation aligns with the fact that degrading a 30 mg/L MO solution is exceptionally challenging, as described above. In stark contrast, an entirely different outcome was observed when a 30 mg/L MO solution, suspended with Al2O3 I, was magnetically stirred in a Si-coated beaker. As illustrated in Figure 2b, the absorption peak at 450 nm completely disappeared after 120 min of magnetic stirring. Concurrently, an obvious color change, from yellow to colorless, was observed for the solution, as indicated in the inset. When Al2O3 I was replaced with Al2O3 II, a similar result was obtained, albeit with slower changes in absorption peak and dye color, as demonstrated in Figure 2c. Conversely, when the solution lacked suspended particles, no observable alterations occurred when the solution was magnetically stirred in a Si-coated beaker, as shown in Figure 2d. Obviously, the friction between Al2O3 and silicon was pivotal in the observed degradation.
However, it is worth pointing out that a new peak at 250 nm in the UV region appeared when the peak at 450 nm disappeared in the UV–Vis adsorption spectra presented in Figure 2b,c. A similar phenomenon had been observed in the context of photocatalytic degradation of MO, where it was suggested that MO molecules were broken into small molecules of some byproducts [34,35,36]. Mass spectrometry tests were performed to identify the byproducts obtained in this study. A 30 mL solution of 30 mg/L MO, suspended with 300 mg of Al2O3 I, was magnetically stirred using a Teflon magnetic rotary disk at 400 rpm in a Si-coated glass beaker at 25 °C in darkness. Samples were extracted separately from the original MO solution, after 90 and 240 min of magnetic stirring, for mass spectrometry analyses. The outcomes of these analyses are presented in Figure 3. In Figure 3a, a prominent mass spectral peak at m/z = 304 is evident for the original MO solution, corresponding to the ionization of the methyl orange parent molecule. As the stirring time reached 90 min, the intensity of the absorption peak at m/z = 304 declined by at least half. Simultaneously, new peaks emerged at m/z = 150, m/z = 122, m/z = 118, and m/z = 109, representing intermediates in the degradation process of methyl orange molecule. After 240 min of stirring, the peak at m/z = 304 disappeared, and the MZ signals of those intermediate products exhibited a slight decrease.
The addition of •OH radicals to the azo double bond has been proposed as the initial step in the oxidative bond cleavages for azo dyes [37,38,39]. This process is believed to account for the variation from m/z = 304 to m/z = 150 (nitro-so-N,N-dimethylaniline) observed in the ion mass spectrums in this study, and the possible destruction process is depicted in Figure 4a. Three major degradation byproducts are displayed in Figure 4b, with m/z = 122 corresponding to benzoic acid [40], m/z = 118 to succinic acid [41], and m/z = 109 to p-phenol [42]. Clearly, the degradation observed in this study is a partial degradation and it remains a challenge to degrade concentrated MO solutions into non-toxic and harmless water and CO2.
It is worth noting that there were no observable changes in either Al2O3 I or Al2O3 II morphology after undergoing extended magnetic stirring for dozens of hours. Since silicon is considerably softer than Al2O3, inspecting the surface of silicon single crystals after friction with Al2O3 during magnetic stirring is imperative. In Figure 5a, an optical microscope image illustrates the surface of an as-received silicon single crystal, which appears exceptionally smooth with minimal defects. Conversely, Figure 5b reveals that some scratches 1–2 μm wide were observed on the surface of a silicon single crystal after being treated as a coating through 10 h of magnetic stirring with Al2O3 I. Despite these scratches, the successful degradation of methyl orange was repeated in a beaker coated with this silicon single crystal, indicating that these surface imperfections had no adverse effect on subsequent catalytic utilization.
In catalytic applications, particular active radicals are formed first, which subsequently drive various specific reactions. The EPR spectra obtained in this study are shown in Figure 6. In a silicon-coated beaker containing a DMPO aqueous solution suspended with Al2O3 I, an unmistakable characteristic peak corresponding to hydroxyl radicals (1:2:2:1) was observed in the EPR measurement after 30 min of magnetic stirring using a Teflon magnetic rotary disk, as shown in Figure 6a [43], while for methanol added with DMPO suspended with Al2O3 I in a silicon-coated beaker, four characteristic peaks representing superoxide radicals appeared in a 1:1:1:1:1 ratio after 30 min of magnetic stirring, as illustrated in Figure 6b [44]. In contrast, no characteristic signals were detected in the two control experiments, with no Al2O3 particles suspended in the solutions. These results suggest that the generation of hydroxyl and superoxide radicals is a consequence of the friction between Al2O3 and silicon in magnetic stirring. Regarding the disparity in MO degradation observed between Al2O3 I and Al2O3 II, it is likely attributed to the fact that Al2O3 I, with its much smaller particles, forms more friction with silicon single crystals compared to Al2O3 II when subjected to the same magnetic stirring conditions.
In tribo-catalytic investigations, it is widely accepted that friction energy excites electron-hole pairs in materials during friction [45,46,47], which subsequently induces redox reactions in ambient environments. Given the degradation of MO solutions associated with silicon single crystals observed in this study, it is reasonable to assume that, through the friction between Al2O3 particles and silicon single crystals in magnetic stirring, as shown in Figure 7, electron-hole pairs are excited in silicon:
Si   F r i c t i o n   e n e r g y   Si + e + h +
Then, the electrons and holes diffuse into the ambient solutions, leading to the formation of some radicals and, ultimately, the degradation of the MO dye:
OH + h +   · OH
O 2 + e   · O 2
· OH   or · O 2 + MO   Dye Decomposition
It is noteworthy that, in previous investigations of tribo-catalytic degradation of organic dyes, electron-hole pairs were consistently excited in particulate semiconductors by friction energy [5,15], which then diffused to the surfaces of the particles, inducing redox reactions in the ambient environment. This study marks the first instance where bulk semiconductors have electron-hole pairs excited through friction energy, instigating redox reactions in the ambient environment.
As previously described, MO solutions with concentrations as high as 30 mg/L have proven challenging to degrade through current catalytic technologies, including photo-catalysis. The findings of this study thus underscore the potential of tribo-catalysis in harnessing mechanical energy for environmental remediation.

4. Conclusions

An effect has been observed in the tribo-catalytic degradation of concentrated MO solutions when utilizing silicon single crystals as coatings. In glass beakers coated with silicon single crystals, 30 mg/L MO solutions suspended with alumina nanoparticles were changed from yellow to colorless within 120 min when Teflon magnetic rotary disks were driven to rotate at 400 rpm. Notably, the characteristic absorption peak of MO at 450 nm gradually weakened and eventually vanished, while a new absorption peak at 250 nm emerged in the UV-Vis spectrum over time. In-depth mass spectrometry tests further illuminated the degradation process, revealing that •OH radicals initiated the breakdown of the nitrogen-nitrogen double bond within MO molecules. This process produced three major intermediate products: benzoic acid (with m/z = 122), succinic acid (m/z = 118), and p-phenol (m/z = 109). Some comparison experiments have been conducted to show that the friction between silicon and alumina in magnetic stirring has resulted in the observed degradation, and hydroxyl and superoxide radicals have been detected to generate from the friction through EPR analysis. It is proposed that electron-hole pairs are excited in silicon single crystals due to the friction with alumina, which diffuse to the surface of the single crystals, resulting in redox reactions in an ambient environment. These results suggest the potential of using silicon in the tribo-catalytic degradation of concentrated MO solutions.

Author Contributions

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

Funding

This research was funded by the National Key R&D Program of China under Grant No. 2023YFE0204400, and by the National Natural Science Foundation of China under Grant No. U2067207.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data underlying this article will be shared on reasonable request to the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ren, X.; Dong, L.; Xu, D.; Hu, B. Challenges towards hydrogen economy in China. Int. J. Hydrogen Energy 2020, 45, 34326–34345. [Google Scholar] [CrossRef]
  2. Goldthau, A.; Tagliapietra, S. Energy crisis: Five questions that must be answered in 2023. Nature 2022, 612, 627–630. [Google Scholar] [CrossRef]
  3. Zor, S. Conservation or revolution? The sustainable transition of textile and apparel firms under the environmental regulation: Evidence from China. J. Clean. Prod. 2023, 382, 135339. [Google Scholar] [CrossRef]
  4. Singha, K.; Pandit, P.; Maity, S.; Sharma, S.R. Harmful environmental effects for textile chemical dyeing practice. In Green Chemistry for Sustainable Textiles; Ibrahim, N., Hussain, C.M., Eds.; Woodhead Publishing: Sawston, UK, 2021; pp. 153–164. [Google Scholar]
  5. 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]
  6. Fan, F.; Xie, S.; Wang, G.; Tian, Z. Tribocatalysis: Challenges and perspectives. Sci. China Chem. 2021, 64, 1609–1613. [Google Scholar] [CrossRef]
  7. Tong, W.; An, Q.; Li, Y.; Li, X.; Zhang, Y. Weak-force energy development and its self-powered environmental purification. Chin. Sci. Bull. 2023, 68, 1381–1391. [Google Scholar] [CrossRef]
  8. 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]
  9. Jin, Z.; Zheng, X.; Zhu, Z.; Hu, C.; Liu, L.; Fang, L.; Luo, N.; Cheng, Z. Enhanced triboelectric degradation of organics by tuning the synergy between ferroelectric polarization and illumination. Mater. Today Chem. 2023, 29, 101427. [Google Scholar] [CrossRef]
  10. Jin, Z.; Zheng, X.; Zhu, Z.; Hu, C.; Liu, L.; Fang, L.; Cheng, Z. Enhanced triboelectric degradation of organics by regulating oxygen vacancies and constructing heterojunctions. Appl. Surf. Sci. 2023, 625, 157228. [Google Scholar] [CrossRef]
  11. Xu, Y.; Yin, R.; Zhang, Y.; Zhou, B.; Sun, P.; Dong, X. Unveiling the mechanism of frictional catalysis in water by Bi12TiO20: A charge transfer and contaminant decomposition path study. Langmuir 2022, 38, 14153–14161. [Google Scholar] [CrossRef]
  12. Yu, H.; Fu, J.; Zhu, X.; Zhao, Z.; Sui, X.; Sun, S.; He, X.; Zhang, Y.; Ye, W. Tribocatalytic degradation of organic pollutants using Fe2O3 nanoparticles. ACS Appl. Nano Mater. 2023, 6, 14364–14373. [Google Scholar] [CrossRef]
  13. Wu, M.; Xu, Y.; He, Q.; Sun, P.; Weng, X.; Dong, X. Tribocatalysis of homogeneous material with multi-size granular distribution for degradation of organic pollutants. J. Colloid Interface Sci. 2022, 622, 602–611. [Google Scholar] [CrossRef] [PubMed]
  14. Zhao, J.; Chen, L.; Luo, W.; Li, H.; Wu, Z.; Xu, Z.; Zhang, Y.; Zhang, H.; Yuan, G.; Gao, J. Strong tribo-catalysis of zinc oxide nanorods via triboelectrically-harvesting friction energy. Ceram. Int. 2020, 46, 25293–25298. [Google Scholar] [CrossRef]
  15. Cui, X.; Li, P.; Lei, H.; Tu, C.; Wang, D.; Wang, Z.; Chen, W. Greatly enhanced tribocatalytic degradation of organic pollutants by TiO2 nanoparticles through efficiently harvesting mechanical energy. Sep. Purif. Technol. 2022, 289, 120814. [Google Scholar] [CrossRef]
  16. Sun, C.; Guo, X.; Ji, R.; Hu, C.; Liu, L.; Fang, L.; Cheng, Z.; Luo, N. Strong tribocatalytic dye degradation by tungsten bronze Ba4Nd2Fe2Nb8O30. Ceram. Int. 2021, 47, 5038–5043. [Google Scholar] [CrossRef]
  17. Sun, C.; Guo, X.; Hu, C.; Liu, L.; Fang, L.; Cheng, Z.; Luo, N. Tribocatalytic degradation of dyes by tungsten bronze ferroelectric Ba2.5Sr2.5Nb8Ta2O30 submicron particles. RSC Adv. 2021, 11, 13386–13395. [Google Scholar] [CrossRef] [PubMed]
  18. Hu, J.; Ma, W.; Pan, Y.; Cheng, Z.; Yu, S.; Gao, J.; Zhang, Z.; Wan, C.; Qiu, C. Insights on the mechanism of Fe doped ZnO for tightly-bound extracellular polymeric substances tribo-catalytic degradation: The role of hydration layers at the interface. Chemosphere 2021, 276, 130170. [Google Scholar] [CrossRef]
  19. Li, P.; Tang, C.; Xiao, X.; Jia, Y.; Chen, W. Flammable gases produced by TiO2 nanoparticles under magnetic stirring in water. Friction 2022, 10, 1127–1133. [Google Scholar] [CrossRef]
  20. Li, P.; Tang, C.; Chen, L.; Hu, Y.; Xiao, X.; Chen, W. Reduction of CO2 by TiO2 nanoparticles through friction in water. Acta Phys. Sin. 2021, 70, 214601. [Google Scholar] [CrossRef]
  21. Yang, B.; Chen, H.; Guo, X.; Wang, L.; Xu, T.; Bian, J.; Yang, Y.; Liu, Q.; Du, Y.; Lou, X. Enhanced tribocatalytic degradation using piezoelectric CdS nanowires for efficient water remediation. J. Mater. Chem. C 2020, 8, 14845–14854. [Google Scholar] [CrossRef]
  22. Zhang, Q.; Jia, Y.; Wang, X.; Zhang, L.; Yuan, G.; Wu, Z. Efficient tribocatalysis of magnetically recyclable cobalt ferrite nanoparticles through harvesting friction energy. Sep. Purif. Technol. 2023, 307, 122846. [Google Scholar] [CrossRef]
  23. Liu, S.; Yang, Y.; Hu, Y.; Rao, W. Effect of strontium substitution on the tribocatalytic performance of barium titanate. Materials 2023, 16, 3160. [Google Scholar] [CrossRef] [PubMed]
  24. Tang, Q.; Zhu, M.; Zhang, H.; Gao, J.; Kwok, K.W.; Kong, L.; Jia, Y.; Liu, L.; Peng, B. Enhanced tribocatalytic degradation of dye pollutants through governing the charge accumulations on the surface of ferroelectric barium zirconium titanate particles. Nano Energy 2022, 100, 107519. [Google Scholar] [CrossRef]
  25. Li, X.; Wang, J.; Zhang, J.; Zhao, C.; Wu, Y.; He, Y. Cadmium sulfide modified zinc oxide heterojunction harvesting ultrasonic mechanical energy for efficient decomposition of dye wastewater. J. Colloid Interface Sci. 2022, 607, 412–422. [Google Scholar] [CrossRef] [PubMed]
  26. Wu, M.; Zhang, Y.; Yi, Y.; Zhou, B.; Sun, P.; Dong, X. Regulation of friction pair to promote conversion of mechanical energy to chemical energy on Bi2WO6 and realization of enhanced tribocatalytic activity to degrade different pollutants. J. Hazard. Mater. 2023, 459, 132147. [Google Scholar] [CrossRef] [PubMed]
  27. Lei, H.; Jia, X.; Wang, H.; Cui, X.; Jia, Y.; Fei, L.; Chen, W. Tribo-catalytic conversions of H2O and CO2 by NiO particles in reactors with plastic and metallic coatings. Coatings 2023, 13, 396. [Google Scholar] [CrossRef]
  28. Jia, X.; Wang, H.; Lei, H.; Mao, C.; Cui, X.; Liu, Y.; Jia, Y.; Yao, W.; Chen, W. Boosting tribo-catalytic conversion of H2O and CO2 by Co3O4 nanoparticles through metallic coatings in reactors. J. Adv. Ceram. 2023. [Google Scholar] [CrossRef]
  29. Cui, X.; Wang, H.; Lei, H.; Jia, X.; Jiang, Y.; Fei, L.; Jia, Y.; Chen, W. Surprising tribo-catalytic conversion of H2O and CO2 into flammablegases utilizing frictions of copper in Water. ChemistrySelect 2023, 8, e202204146. [Google Scholar] [CrossRef]
  30. Li, X.; Tong, W.; Shi, J.; Chen, Y.; Zhang, Y.; An, Q. Tribocatalysis mechanisms: Electron transfer and transition. J. Mater. Chem. A 2023, 11, 4458–4472. [Google Scholar] [CrossRef]
  31. Zhai, L.; Bai, Z.; Zhu, Y.; Wang, B.; Luo, W. Fabrication of chitosan microspheres for efficient adsorption of methyl orange. Chin. J. Chem. Eng. 2018, 26, 657–666. [Google Scholar] [CrossRef]
  32. Sanromán, M.Á.; Pazos, M.; Cameselle, C. Optimisation of electrochemical decolourisation process of an azo dye, methyl orange. J. Chem. Technol. Biotechnol. 2004, 79, 1349–1353. [Google Scholar] [CrossRef]
  33. Liu, T.; Wang, L.; Lu, X.; Fan, J.; Cai, X.; Gao, B.; Miao, R.; Wang, J.; Lv, Y. Comparative study of the photocatalytic performance for the degradation of different dyes by ZnIn2S4: Adsorption, active species, and pathways. RSC Adv. 2017, 7, 12292–12300. [Google Scholar] [CrossRef]
  34. Li, X.; Zhu, J.; Li, H. Comparative study on the mechanism in photocatalytic degradation of different-type organic dyes on SnS2 and CdS. Appl. Catal. B Environ. 2012, 123, 174–181. [Google Scholar] [CrossRef]
  35. Wang, M.; Li, M.; Xu, L.; Wang, L.; Ju, Z.; Li, G.; Qian, Y. High yield synthesis of novel boron nitride submicro-boxes and their photocatalytic application under visible light irradiation. Catal. Sci. Technol. 2011, 1, 1159–1165. [Google Scholar] [CrossRef]
  36. Filice, S.; D’Angelo, D.; Libertino, S.; Nicotera, I.; Kosma, V.; Privitera, V.; Scalese, S. Graphene oxide and titania hybrid nafion membranes for efficient removal of methyl orange dye from water. Carbon 2015, 82, 489–499. [Google Scholar] [CrossRef]
  37. Joseph, J.M.; Destaillats, H.; Hung, H.M.; Hoffmann, M.R. The sonochemical degradation of azobenzene and related azo dyes: Rate enhancements via Fenton’s reactions. J. Phys. Chem. A 2000, 104, 301–307. [Google Scholar] [CrossRef]
  38. Wojnárovits, L.; Takács, E. Irradiation treatment of azo dye containing wastewater: An overview. Radiat. Phys. Sonochem. 2008, 77, 225–244. [Google Scholar] [CrossRef]
  39. Okitsu, K.; Iwasaki, K.; Yobiko, Y.; Bandow, H.; Nishimura, R.; Maeda, Y. Sonochemical degradation of azo dyes in aqueous solution: A new heterogeneous kinetics model taking into account the local concentration of OH radicals and azo dyes. Ultrason. Sonochem. 2005, 12, 255–262. [Google Scholar] [CrossRef]
  40. Ariyanti, D.; Maillot, M.; Gao, W. Photo-assisted degradation of dyes in a binary system using TiO2 under simulated solar radiation. J. Environ. Chem. Eng. 2018, 6, 539–548. [Google Scholar] [CrossRef]
  41. Panda, N.; Sahoo, H.; Mohapatra, S. Decolourization of methyl orange using Fenton-like mesoporous Fe2O3–SiO2 composite. J. Hazard. Mater. 2011, 185, 359–365. [Google Scholar] [CrossRef]
  42. Khan, M.A.; Mutahir, S.; Wang, F.; Lei, W.; Xia, M.; Zhu, S. Facile one-step economical methodology of metal free g-C3N4 synthesis with remarkable photocatalytic performance under visible light to degrade trans-resveratrol. J. Hazard. Mater. 2019, 367, 293–303. [Google Scholar] [CrossRef]
  43. Lin, F.; Zhang, Y.; Wang, L.; Zhang, Y.; Wang, D.; Yang, M.; Yang, J.; Zhang, B.; Jiang, Z.; Li, C. Highly efficient photocatalytic oxidation of sulfur-containing organic compounds and dyes on TiO2 with dual cocatalysts Pt and RuO2. Appl. Cata. B Environ. 2012, 127, 363–370. [Google Scholar] [CrossRef]
  44. Duan, Y.; Luo, J.; Zhou, S.; Mao, X.; Shah, M.W.; Wang, F.; Chen, Z.; Wang, C. TiO2-supported Ag nanoclusters with enhanced visible light activity for the photocatalytic removal of NO. Appl. Catal. B Environ. 2018, 234, 206–212. [Google Scholar] [CrossRef]
  45. Geng, L.; Qian, Y.; Song, W.; Bao, L. Enhanced tribocatalytic pollutant degradation through tuning oxygen vacancy in BaTiO3 nanoparticles. Appl. Surf. Sci. 2023, 637, 157960. [Google Scholar] [CrossRef]
  46. Schwab, T.; Thomele, D.; Aicher, K.; Dunlop, J.W.; McKenna, K.; Diwald, O. Rubbing powders: Direct spectroscopic observation of triboinduced oxygen radical formation in MgO nanocube ensembles. J. Phys. Chem. C 2021, 125, 22239–22248. [Google Scholar] [CrossRef] [PubMed]
  47. Alabbad, E.A.; Bashir, S.; Liu, J.L. Efficient removal of direct yellow dye using chitosan crosslinked isovanillin derivative biopolymer utilizing triboelectric energy produced from homogeneous catalysis. Catal. Today 2022, 400, 132–145. [Google Scholar] [CrossRef]
Figure 1. SEM images of Al2O3 powders used in this study: (a) α-Al2O3 nanoparticles, termed as Al2O3 I; (b) α-Al2O3 laminated powder, termed as Al2O3 II.
Figure 1. SEM images of Al2O3 powders used in this study: (a) α-Al2O3 nanoparticles, termed as Al2O3 I; (b) α-Al2O3 laminated powder, termed as Al2O3 II.
Coatings 13 01804 g001
Figure 2. UV–Vis adsorption spectra and color evolution for a 30 mg/L MO solution in the course of magnetic stirring: (a) suspended with Al2O3 I in a glass beaker; (b) suspended with Al2O3 I in a Si-coated beaker; (c) suspended with Al2O3 II in a Si-coated beaker; (d) with no particles suspended in a Si-coated beaker.
Figure 2. UV–Vis adsorption spectra and color evolution for a 30 mg/L MO solution in the course of magnetic stirring: (a) suspended with Al2O3 I in a glass beaker; (b) suspended with Al2O3 I in a Si-coated beaker; (c) suspended with Al2O3 II in a Si-coated beaker; (d) with no particles suspended in a Si-coated beaker.
Coatings 13 01804 g002
Figure 3. Mass spectrums: (a) original 30 mg/L MO solution; (b) after 90 min of magnetic stirring; (c) after 240 min of magnetic stirring with Al2O3 I in a Si-coated beaker.
Figure 3. Mass spectrums: (a) original 30 mg/L MO solution; (b) after 90 min of magnetic stirring; (c) after 240 min of magnetic stirring with Al2O3 I in a Si-coated beaker.
Coatings 13 01804 g003
Figure 4. (a) Attack of •OH radicals to the azo double bond; (b) Structural formulas for three main byproducts, to which methyl orange is converted.
Figure 4. (a) Attack of •OH radicals to the azo double bond; (b) Structural formulas for three main byproducts, to which methyl orange is converted.
Coatings 13 01804 g004
Figure 5. Optical microscope image for the surface of: (a) an as-received single crystal silicon; (b) a silicon single crystal after being treated as a coating through magnetic stirring with Al2O3 I for 10 h.
Figure 5. Optical microscope image for the surface of: (a) an as-received single crystal silicon; (b) a silicon single crystal after being treated as a coating through magnetic stirring with Al2O3 I for 10 h.
Coatings 13 01804 g005
Figure 6. EPR spectra with DMPO as a spin-trap reagent were obtained for deionized water and methanol solution, where Al2O3 nanoparticles were magnetically stirred using a Teflon magnetic rotary disk separately in a Si-coated beaker: (a) in deionized water, detecting hydroxyl radicals; (b) in methanol solution, detecting superoxide radicals.
Figure 6. EPR spectra with DMPO as a spin-trap reagent were obtained for deionized water and methanol solution, where Al2O3 nanoparticles were magnetically stirred using a Teflon magnetic rotary disk separately in a Si-coated beaker: (a) in deionized water, detecting hydroxyl radicals; (b) in methanol solution, detecting superoxide radicals.
Coatings 13 01804 g006
Figure 7. Mechanism diagram for the excitation of electron-hole pairs in silicon single crystals through the friction with Al2O3 particles in magnetic stirring.
Figure 7. Mechanism diagram for the excitation of electron-hole pairs in silicon single crystals through the friction with Al2O3 particles in magnetic stirring.
Coatings 13 01804 g007
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Cui, X.; Guo, Z.; Lei, H.; Jia, X.; Mao, C.; Ruan, L.; Zhou, X.; Wang, Z.; Chen, F.; Chen, W. Tribo-Catalytic Degradation of Methyl Orange Solutions Enhanced by Silicon Single Crystals. Coatings 2023, 13, 1804. https://doi.org/10.3390/coatings13101804

AMA Style

Cui X, Guo Z, Lei H, Jia X, Mao C, Ruan L, Zhou X, Wang Z, Chen F, Chen W. Tribo-Catalytic Degradation of Methyl Orange Solutions Enhanced by Silicon Single Crystals. Coatings. 2023; 13(10):1804. https://doi.org/10.3390/coatings13101804

Chicago/Turabian Style

Cui, Xiaodong, Zhiyu Guo, Hua Lei, Xuchao Jia, Chenyue Mao, Lujie Ruan, Xiaoyuan Zhou, Zhu Wang, Feng Chen, and Wanping Chen. 2023. "Tribo-Catalytic Degradation of Methyl Orange Solutions Enhanced by Silicon Single Crystals" Coatings 13, no. 10: 1804. https://doi.org/10.3390/coatings13101804

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