Next Article in Journal
Automated Quantum Dots Purification via Solid Phase Extraction
Next Article in Special Issue
FeNi LDH/V2CTx/NF as Self-Supported Bifunctional Electrocatalyst for Highly Effective Overall Water Splitting
Previous Article in Journal
Synthesis, Phase Transformations and Strength Properties of Nanostructured (1 − x)ZrO2 − xCeO2 Composite Ceramics
Previous Article in Special Issue
In Situ Construction of ZIF-67-Derived Hybrid Tricobalt Tetraoxide@Carbon for Supercapacitor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Strong Tribocatalytic Nitrogen Fixation of Graphite Carbon Nitride g-C3N4 through Harvesting Friction Energy

1
Xi’an Key Laboratory of Textile Chemical Engineering Auxiliaries, School of Environmental and Chemical Engineering, Xi’an Polytechnic University, Xi’an 710048, China
2
College of Physics and Electronic Information Engineering, Zhejiang Normal University, Jinhua 321004, China
3
Key Laboratory of Surface & Interface Science of Polymer Materials of Zhejiang Province, Department of Chemistry, Zhejiang Sci-Tech University, Hangzhou 310018, China
4
School of Science, Xi’an University of Posts and Telecommunications, Xi’an 710121, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(12), 1981; https://doi.org/10.3390/nano12121981
Submission received: 27 April 2022 / Revised: 7 June 2022 / Accepted: 8 June 2022 / Published: 9 June 2022
(This article belongs to the Special Issue Nanostructure-Based Energy Electrocatalysis)

Abstract

:
Mechanical energy derived from friction is a kind of clean energy which is ubiquitous in nature. In this research, two-dimensional graphite carbon nitride (g-C3N4) is successfully applied to the conversion of nitrogen (N2) fixation through collecting the mechanical energy generated from the friction between a g-C3N4 catalyst and a stirring rod. At the stirring speed of 1000 r/min, the tribocatalytic ammonia radical (NH4+) generation rate of g-C3N4 can achieve 100.56 μmol·L−1·g−1·h−1 using methanol as a positive charge scavenger, which is 3.91 times higher than that without any scavengers. Meanwhile, ammonia is not generated without a catalyst or contact between the g-C3N4 catalyst and the stirring rod. The tribocatalytic effect originates from the friction between the g-C3N4 catalyst and the stirring rod which results in the charges transfer crossing the contact interface, then the positive and negative charges remain on the catalyst and the stirring rod respectively, which can further react with the substance dissolved in the reaction solution to achieve the conversion of N2 to ammonia. The effects of number and stirring speed of the rods on the performance of g-C3N4 tribocatalytic N2 fixation are further investigated. This excellent and efficient tribocatalysis can provide a potential avenue towards harvesting the mechanical energy in a natural environment.

1. Introduction

Nowadays, due to the immoderate consumption and mining of fossil fuels, energy shortage and environment pollution have become critical issues which are a threat to the survival and development of society [1]. Accordingly, it is necessary to look for renewable and green energy to replace these fossil fuels. Ammonia has been regarded as a green energy source with some advantages such as no carbon dioxide emission, high energy density and convenient transportation [2,3]. Nevertheless, how to perform ammonia production is also a vital issue. At present, various methods have been reported to convert nitrogen (N2) to ammonia (NH3), such as thermocatalytic reduction, electrocatalytic reduction and photocatalytic reduction [4,5,6]. However, thermocatalytic and electrocatalytic reduction usually require high-pressure or high-temperature operating conditions, which limit the actual application of ammonia production [7,8]. Additionally, photocatalytic reduction of nitrogen can be performed at a mild condition, but low light utilization efficiency would seriously hinder the actual production of ammonia [9]. Therefore, it is crucial to seek mild and highly efficient approaches for nitrogen reduction.
In nature, mechanical energy is widespread distributed energy, which exists in wind, river flows and human movement [10]. If such abundant energy can be successfully harvested for the reduction of nitrogen, it would be meaningful to propel the production of ammonia. Tribocatalysis, which can convert the friction of mechanical energy into the electric energy via persistent friction, has attracted the attention of researchers [11,12,13]. Under friction, when two different kinds of materials contact each other, chemical bonds will form through physical contact on the interface [14]. After being separated, two materials will carry the positive or negative charges respectively, due to breakage of the chemical bonds [15]. Then the free charges generated via the triboelectric process can be further applied to the catalytic reaction such as dye degradation, carbon dioxide or flammable gas production [16,17,18,19,20]. However, to date, there has been no report about harvesting the mechanical energy from friction via tribocatalysis for nitrogen reduction to produce ammonia.
Graphite carbon nitride (g-C3N4) is an emerging semiconductor material with layered structure similar to graphite [21]. The C and N atoms inside it are arranged alternately through sp2 hybridization [22]. g-C3N4 is a candidate catalyst in several catalytic areas such as the decomposition of organic pollutants [23,24,25], hydrogen evolution [26,27] and carbon dioxide reduction [28,29,30] based on its advantages of being insoluble in water and having a large specific surface area and stable chemical properties. In 2019, Xia et al. achieved the photocatalytic synthesis of ammonia from nitrogen through using a defecting g-C3N4 catalyst. After 100 min light irradiation, the yield of ammonia can be up to 54 μmol/L [31]. Dong et al. achieved highly efficient dichlorophenols decomposition via photocatalysis using g-C3N4 [23]. Recent studies have found that mesoporous carbon nitride materials can exhibit excellent catalytic activity through introducing metal atoms inside them. Gianvito Vilé et al. have reported that Cu-based single-atom catalysts developed on a mesoporous carbon nitride carrier exhibited excellent catalytic activity during the synthesis of triazoles [32]. Liu et al. found that the photocatalytic decomposition activity of gemfibrozil can be significantly improved by introducing Ag or Cu atoms into the mesoporous carbon nitride skeleton, which is related to ligand-to-metal charge transfer (LMCT) or ligand-to-metal-to-ligand charge transfer (LMLCT) [33]. Based on the above analysis, g-C3N4 is hopeful for applications in highly efficient tribocatalytic nitrogen fixation, which is not reported at present.
In this work, the excellent tribocatalytic reduction of nitrogen to ammonia under stirring is achieved in g-C3N4 which is fabricated via the chemical blowing method. After 10 h stirring at room temperature in the dark, the tribocatalytic ammonia radical (NH4+) generation rate is up to about 100.56 μmol·L−1·g−1·h−1 using a positive charge scavenger (methanol), which is 3.91 times higher than that without any charge scavengers. In addition, the effects of a negative charge scavenger, the number and speed of stirring rods, and the contact area between the catalyst and stirring rods in tribocatalysis on the performance of g-C3N4 tribocatalytic N2 fixation are also investigated. The possible tribocatalytic mechanism of N2 fixation has been also proposed. As the N2 fixation research continues to thrive and expand, the finding in this work provides a great potential application to harvest the mechanical energy via tribocatalysis for clean energy production.

2. Materials and Methods

2.1. Preparation of g-C3N4 Sample

g-C3N4 used in this research was prepared according to the reported chemical blowing method [34,35]. A certain amount of ammonium chloride (16 g) and melamine (4 g) were accurately weighed and then mixed together. The mixture was thoroughly ground in a mortar and placed in a crucible. Then, it was stuffed into the muffle furnace, heated from room temperature to 550 °C (6 °C/min), and the calcination time was set to 4 h. After the calcination, the faint yellow product in the crucible was collected and ground through an agate mortar to obtain g-C3N4.

2.2. Characterization

The X-ray diffraction pattern of the g-C3N4 sample was tested on a D8 Advance diffractometer (Bruker AXS, Karlsruhe, Germany). The micro morphology of the sample was examined with a scanning electron microscope (Gemini SEM 300, ZEISS, Oberkochen, Germany). The chemical properties of the sample were analyzed through using the X-ray photoelectron spectrometer (XPS, ESCALAB 250Xi, Waltham, MA, USA). The infrared spectra of the sample prepared through the KBr tablet pressing method were characterized via the Fourier transform infrared spectrometer (FTIR, Nicolet NEXUS 670, Ramsey, MN, USA). The content of ammonia (NH4+) was analyzed via a UV-Vis spectrophotometer (Ocean Optics QE65Pro, Dunedin, FL, USA).

2.3. Tribocatalytic Performance Measurements

To investigate g-C3N4 tribocatalytic performance, N2 fixation experiments were performed. A total of 50 mL methanol solution (containing 5 mL methanol and 45 mL DI water) mixed with 50 mg g-C3N4 was contained in a brown bottle. The solution was then placed under shading conditions and stirred at 1000 rpm for 2 h through a polytetrafluoroethylene (PTFE) stirring rod with a specification of 6 × 20 mm2 to achieve adsorption-desorption equilibrium. Then under continuous stirring, the suspension of 3 mL was collected through a rubber-tipped dropper every 2 h of stirring. The supernatan was obtained through a centrifuge (3 min, 6500 rpm). The generation of NH4+ produced in the process of tribocatalysis was determined through the colorimetric method, and Nessler reagent was selected as the indicator [36,37]. Then 40 μL sodium tartrate solution (0.5 g·mL−1) and 60 μL Nessler’s reagent were added dropwise into supernatant and left to stand for 12 min to react sufficiently. Finally, the content of ammonium was analyzed at the peak of ~420 nm via a UV–Vis spectrophotometer.

3. Results and Discussion

SEM images of the g-C3N4 sample before the tribocatalytic N2 fixation reaction are depicted in Figure 1. g-C3N4 samples show the agglomerate morphology composed of many irregular ultra-thin two-dimensional sheet-like structures. From Figure 1, g-C3N4 catalyst material composed of many huge flakes has a large specific surface area.
The crystal diffraction patterns of g-C3N4 have been measured with XRD, as shown in Figure 2. The obvious peaks at 2θ value of 12.93° and 27.69° are corresponding to the (100) and (002) crystal plane of g-C3N4 through referring the standard card PDF#87-1526. It is ascribed to the orderly stacking of the conjugated carbon-nitrogen heterocycles in a planar and layered framework, respectively [38,39]. In addition, the high-intensity diffraction peaks and the absence of other impurity peaks indict the excellent synthesis of g-C3N4.
FTIR spectra depicted in Figure 3 reveal the functional groups of g-C3N4 before the tribocatalytic N2 fixation. The peak around the wave number of 812 cm−1 is caused by the out-of-plane bending vibration of the triazine structure [40,41]. The peaks in the wave number range of 1240–1640 cm−1 are related to the stretching vibration mode of C–N heterocycle in the g-C3N4 sample [23]. The broad peaks in the range of 3080–3450 cm−1 are ascribed to the stretching vibrations of N−H and O−H groups [39,42].
The elemental states of the g-C3N4 sample have been performed with XPS measurement, as depicted in Figure 4. From Figure 4a, the measured survey spectrum confirms the inclusion of both C and N elements in g-C3N4. The high-resolution XPS spectrum of C 1s are presented in Figure 4b. The spectrum of C 1s has two peaks at 284.42 eV and 287.22 eV, which are assign to graphitic carbon adsorbed on the sample surface and sp2-bonded carbon in the triazine structure [39]. As shown in Figure 4c, the N 1s spectrum of g-C3N4 was deconvoluted into three peaks at binding energies of 397.72 eV, 398.87 eV and 400.07 eV, corresponding to C=N−C, N−(C)3 and N−H in the sample, respectively [38,43].
To investigate the tribocatalytic activity of the g-C3N4 sample, N2 fixation experiments were performed under stirring. Figure 5 depicts the tribocatalytic N2 fixation performance with the different scavengers. When methanol is added as the positive charge scavenger, with the increase of the stirring time, the NH4+ content increases linearly [44]. The generation rate of NH4+ can reach 100.56 μmol·L−1·g−1·h−1 after 10 h stirring. It is 3.91 times higher than that without any scavengers. Since the methanol can consume the positive charges, the generation of the negative charges is promoted, which can enhance the tribocatalytic N2 fixation performance effectively. It is worth noting that the reductive active radicals, such as the negative charges, are necessary in the N2 fixation reaction [45]. TBA can react with hydroxyl radicals (OH) in solution as the radical scavenger [46,47]. Since the formation of ·OH requires the participation of positive charge (q+) and hydroxide ion (OH), the addition of TBA promotes the consumption of q+ and effectively prolongs the lifetime of negative charge (q). Therefore, the addition of TBA is beneficial to the tribocatalytic nitrogen fixation of g-C3N4. Consequently, KBrO3 as the negative charge scavenger is used to investigate the important role of the negative charges in this catalytic process [48]. It can be observed that the NH4+ generation rate is about 0.26 μmol·L−1·g−1·h−1. It can be concluded that the addition of KBrO3 is not beneficial for the N2 fixation reaction, which accords with the theoretic expectation.
To further investigate the schematic mechanism of the tribocatalytic N2 fixation reaction, the control experiments under the different addition have been performed as shown in Figure 6. There is scarcely any generation of NH4+ without a catalyst, which indicates that the addition of a catalyst is essential for the tribocatalytic N2 fixation reaction. Additionally, the tribocatalytic performance associates with the contact area between the g-C3N4 catalyst and the stirring rod [49]. Therefore, control experiments with the different stirring rods were carried out. Rod I is the commercial PTFE-sealed rod. Two polyvinyl chloride (PVC) electrical tape rings with a width of 1mm were wound on the stirring rod to avoid contact between the catalyst and stirring rod as far as possible, and the stirring rod is named rod II. Obviously, with the decrease of the contact area, the NH4+ generation rate reduces gradually, indicating that the contact area is an important influencing factor of the tribocatalytic performance.
The influence of the stirring speed on the tribocatalytic performance was also considered. The much faster stirring speed provides much more contact times per minute, that is to say, the contact area is also enlarged relatively per minute. Consequently, the tribocatalytic performance would be enhanced. As depicted in Figure 7, the NH4+ generation rate is improved significantly from 15.63 to 100.56 μmol·L−1·g−1·h−1 with the stirring speed increasing from 400 to 1000 rpm, and a linear relationship between g-C3N4 tribocatalytic nitrogen fixation rate and stirring speed is observed. As the stirring speed increases, the friction frequency increases, and more active substances can be produced to participate in the nitrogen fixation process, so an efficient nitrogen fixation can be obtained.
The effect of contact area on the performance of g-C3N4 tribocatalytic N2 fixation is investigated through adjusting the number of stirring rods. Figure 8 shows the nitrogen fixation effect with a different number of stirring rods. Obviously, the total contact area is strongly enlarged as the number of the stirring rods increases. As expected, the NH4+ generation rate is about 244.02 μmol·L−1·g−1·h−1 using three stirring rods, which is 2.43 times that of only one stirring rod. Typically, the contact area is usually proportional to the number of rods, but the NH4+ generation rate is not. For instance, the distribution of the catalyst is not uniform in suspension, which leads to inadequate contact between the rods and catalyst, that is to say, each stirring time may not necessarily induce the tribocatalytic reaction. Perhaps there are other influencing factors which influence the tribocatalytic performance. Consequently, the NH4+ generation rate is not in linear correlation with the number of stirring rods.
Furthermore, the tribocatalytic nitrogen fixation performance of g-C3N4 was evaluated through comparison with other existing nitrogen fixation research. It can be seen from Table 1 that g-C3N4 can realize nitrogen fixation through both photocatalysis and tribocatalysis, and it has an excellent performance [50]. In addition, due to its large specific surface area, g-C3N4 has superior nitrogen fixation activity compared to other materials [51,52,53,54,55].
According to the previous analysis, the tribocatalytic mechanism of g-C3N4 is drawn in Figure 9. The friction between the g-C3N4 catalyst and the stirring rod is realized via mechanical stirring, which is accompanied by charge transfer. It is known from the empirical rule of triboelectrification, the g-C3N4 catalyst is more likely to lose electrons with positive charge than the stirring rod. Therefore, in the process of tribocatalysis, the stirring rod is negatively charged and the catalyst is positively charged. The above process can be represented by Equation (1) [56]:
g C 3 N 4 + stirring   rod Friction g C 3 N 4 ( q + ) + stirring   rod ( q )
Then methanol as the positive charge scavenger will react with q+ to avoid NH4+ being oxidized [37]. In addition, q will react with N2 to produce NH4+, which is dissolved in DI water to form NH4+. The main reaction process is described in Equations (2)–(4) [57,58]:
q + + CH 3 OH CH 3 OH +
6   q + N 2 + 6 H + 2 NH 3
NH 3 + H 2 O NH 3 · H 2 O NH 4 + + OH
In this work, g-C3N4 exhibits excellent tribocatalytic performance in the N2 fixation process. Though, as a matter of fact, the applications of the tribocatalysis are not restricted in this field. In the past few years, He et al. have achieved excellent tribocatalytic performance on removal of the heavy metal ion Cr6+ through using commercial iron turning with amorphous iron oxides. After 36 h stirring, the removal ratio of Cr6+ can reach about 98% [59]. Additionally, Li et al. have successfully produced flammable gases such as CO, CH4 and H2 through harvesting mechanical energy with TiO2 nanoparticles [20]. In the future, g-C3N4 has potential applications in wastewater treatment and energy generation fields via tribocatalysis.

4. Conclusions

In summary, g-C3N4 has been fabricated successfully via the chemical blowing method and shows excellent tribocatalytic performance in the reduction of nitrogen to NH4+. After 10 h of continuous stirring at 1000 rpm in the dark, the generation rate of NH4+ can reach 100.56 μmol·L−1·g−1·h−1 using methanol as a positive charges scavenger, which is 3.91 times higher than that without any scavengers. Furthermore, the performance of the tribocatalytic nitrogen fixation of g-C3N4 can be effectively optimized through increasing the stirring speed or number of stirring rods. Consequently, g-C3N4 has the remarkable potential application in the tribocatalytic N2 fixation reaction. Tribocatalysis has a bright application prospect in energy development fields such as nitrogen fixation, carbon dioxide reduction and water decomposition in the future.

Author Contributions

Conceptualization, J.G., T.X. and Z.W.; methodology, J.G., L.R. and X.D.; software, J.G.; validation, J.G. and L.R.; formal analysis, Z.W.; investigation, J.G. and L.R.; resources, H.L. and S.H.; data curation, J.G.; writing—original draft preparation, J.G. and L.R.; writing—review and editing, Z.W.; visualization, Z.W. and Y.J.; supervision, Z.W.; 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, 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); Hebei Key Laboratory of Dielectric and Electrolyte Functional Material, Northeastern University at Qinhuangdao (No. HKDEFM2021101); Doctoral Research Start-up Fund project of Xi’an Polytechnic University (Grant Number: 107020589); Shaanxi Provincial High-Level Talents Introduction Project (Youth Talent Fund); Key Research and Development Program of Zhejiang Province (2021C01006).

Data Availability Statement

The data presented in this study are available in this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. 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]
  2. Iwamoto, M.; Akiyama, M.; Aihara, K.; Deguchi, T. Ammonia synthesis on wool-like Au, Pt, Pd, Ag, or Cu electrode catalysts in nonthermal atmospheric-pressure plasma of N2 and H2. ACS Catal. 2017, 7, 6924–6929. [Google Scholar] [CrossRef]
  3. Singh, A.R.; Rohr, B.A.; Schwalbe, J.A.; Cargnello, M.; Chan, K.; Jaramillo, T.F.; Chorkendorff, I.; Nørskov, J.K. Electrochemical Ammonia synthesis—The selectivity challenge. ACS Catal. 2017, 7, 706–709. [Google Scholar] [CrossRef] [Green Version]
  4. Vojvodic, A.; Medford, A.J.; Studt, F.; Abild Pedersen, F.; Khan, T.S.; Bligaard, T.; Nørskov, J.K. Exploring the limits: A low-pressure, low-temperature haber–bosch process. Chem. Phys. Lett. 2014, 598, 108–112. [Google Scholar] [CrossRef]
  5. Sadeghzadeh-Attar, A. Photocatalytic degradation evaluation of N-Fe co-doped aligned TiO2 nanorods based on the effect of annealing temperature. J. Adv. Ceram. 2020, 9, 107–122. [Google Scholar] [CrossRef] [Green Version]
  6. Chen, X.; Li, N.; Kong, Z.; Ong, W.J.; Zhao, X. Photocatalytic fixation of nitrogen to ammonia: State-of-the-art advancements and future prospects. Mater. Horiz. 2018, 5, 9–27. [Google Scholar] [CrossRef]
  7. Luo, Y.; Pu, T.; Fan, S.; Liu, H.; Zhu, J. Enhanced piezoelectric properties in low-temperature sintering PZN-PZT ceramics by adjusting Zr/Ti ratio. J. Adv. Dielect. 2022, 12, 2250001. [Google Scholar] [CrossRef]
  8. Wang, Q.; Guo, J.; Chen, P. Recent progress towards mild-condition ammonia synthesis. J. Energy Chem. 2019, 36, 25–36. [Google Scholar] [CrossRef] [Green Version]
  9. 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]
  10. Panda, P.K.; Sahoo, B.; Sureshkumar, V.; Politova, E.D. Effect of Zr4+ on piezoelectric, dielectric and ferroelectric properties of barium calcium titanate lead-free ceramics. J. Adv. Dielect. 2021, 11, 2150024. [Google Scholar] [CrossRef]
  11. Kajdas, C.; Hiratsuka, K. Tribochemistry, tribocatalysis, and the negative-ion-radical action mechanism. Proc. IMechE Part J J. Eng. Tribol. 2009, 223, 827–848. [Google Scholar] [CrossRef]
  12. Fan, F.R.; Tang, W.; Wang, Z.L. Flexible nanogenerators for energy harvesting and self-powered electronics. Adv. Mater. 2016, 28, 4283–4305. [Google Scholar] [CrossRef] [PubMed]
  13. Fan, F.R.; Xie, S.; Wang, G.W.; Tian, Z.Q. Tribocatalysis: Challenges and perspectives. Sci. China Chem. 2021, 64, 1609–1613. [Google Scholar] [CrossRef]
  14. Wang, Z.L. Triboelectric nanogenerators as new energy technology for self-powered systems and as active mechanical and chemical sensors. ACS Nano 2013, 7, 9533–9557. [Google Scholar] [CrossRef]
  15. Shaw, P.E.; Barton, E.H. Experiments on tribo-electricity. I.—The tribo-electric series. Proc. R. Soc. Lond. A 1917, 94, 16–33. [Google Scholar]
  16. Chen, C.; Wang, Y.; Li, J.; Wu, C.; Yang, G. Piezoelectric, ferroelectric and pyroelectric properties of (100 − x) Pb (Mg1/3Nb2/3) O3 − xPbTiO3 ceramics. J. Adv. Dielect. 2022, 2250002. [Google Scholar] [CrossRef]
  17. Wu, M.; Lei, H.; Chen, J.; Dong, X. Friction energy harvesting on bismuth tungstate catalyst for tribocatalytic degradation of organic pollutants. J. Colloid Interface Sci. 2021, 587, 883–890. [Google Scholar] [CrossRef] [PubMed]
  18. Hu, J.; Ma, W.; Pan, Y.; Chen, Z.; Zhang, Z.; Wan, C.; Sun, Y.; Qiu, C. Resolving the tribo-catalytic reaction mechanism for biochar regulated zinc oxide and its application in protein transformation. J. Colloid Interface Sci. 2022, 607, 1908–1918. [Google Scholar] [CrossRef] [PubMed]
  19. Wang, C. Piezo-catalytic degradation of havriliak–negami type. J. Adv. Dielect. 2019, 9, 1950021. [Google Scholar] [CrossRef]
  20. 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]
  21. Yan, H.; Li, J.; Zhang, M.; Zhao, Y.; Feng, Y.; Zhang, Y. Enhanced corrosion resistance and adhesion of epoxy coating by two-dimensional graphite-like g-C3N4 nanosheets. J. Colloid Interface Sci. 2020, 579, 152–161. [Google Scholar] [CrossRef]
  22. Mishra, A.; Mehta, A.; Basu, S.; Shetti, N.P.; Reddy, K.R.; Aminabhavi, T.M. Graphitic carbon nitride (g–C3N4)–based metal-free photocatalysts for water splitting: A review. Carbon 2019, 149, 693–721. [Google Scholar] [CrossRef]
  23. Lei, H.; Wu, M.; Mo, F.; Ji, S.; Dong, X.; Jia, Y.; Wang, F.; Wu, Z. Efficiently harvesting the ultrasonic vibration energy of two-dimensional graphitic carbon nitride for piezocatalytic degradation of dichlorophenols. Environ. Sci. Nano 2021, 8, 1398–1407. [Google Scholar] [CrossRef]
  24. Yang, W.; Chen, Y.; Gao, S.; Sang, L.; Tao, R.; Sun, C.; Shang, J.K.; Li, Q. Post-illumination activity of Bi2WO6 in the dark from the photocatalytic “memory” effect. J. Adv. Ceram. 2021, 10, 355–367. [Google Scholar] [CrossRef]
  25. Mo, F.; Liu, Y.; Xu, Y.; He, Q.; Sun, P.; Dong, X. Photocatalytic elimination of moxifloxacin by two-dimensional graphitic carbon nitride nanosheets: Enhanced activity, degradation mechanism and potential practical application. Sep. Purif. Technol. 2022, 292, 121067. [Google Scholar] [CrossRef]
  26. Yu, Z.; Mao, K.; Feng, Y. Single-source-precursor synthesis of porous W-containing SiC-based nanocomposites as hydrogen evolution reaction electrocatalysts. J. Adv. Ceram. 2021, 10, 1338–1349. [Google Scholar] [CrossRef]
  27. Han, C.; Su, P.; Tan, B.; Ma, X.; Lv, H.; Huang, C.; Wang, P.; Tong, Z.; Li, G.; Huang, Y.; et al. Defective ultra-thin two-dimensional g-C3N4 photocatalyst for enhanced photocatalytic H2 evolution activity. J. Colloid Interface Sci. 2021, 581, 159–166. [Google Scholar] [CrossRef]
  28. Aggarwal, M.; Basu, S.; Shetti, N.P.; Nadagouda, M.N.; Kwon, E.E.; Park, Y.K.; Aminabhavi, T.M. Photocatalytic carbon dioxide reduction: Exploring the role of ultrathin 2D graphitic carbon nitride (g-C3N4). Chem. Eng. J. 2021, 425, 131402. [Google Scholar] [CrossRef]
  29. Zhang, X.; Yuan, X.; Jiang, L.; Zhang, J.; Yu, H.; Wang, H.; Zeng, G. Powerful combination of 2D g-C3N4 and 2D nanomaterials for photocatalysis: Recent advances. Chem. Eng. J. 2020, 390, 124475. [Google Scholar] [CrossRef]
  30. Sun, Z.; Wang, H.; Wu, Z.; Wang, L. g-C3N4 based composite photocatalysts for photocatalytic CO2 reduction. Catal. Today 2018, 300, 160–172. [Google Scholar] [CrossRef]
  31. Zhang, Y.; Di, J.; Ding, P.; Zhao, J.; Gu, K.; Chen, X.; Yan, C.; Yin, S.; Xia, J.; Li, H. Ultrathin g-C3N4 with enriched surface carbon vacancies enables highly efficient photocatalytic nitrogen fixation. J. Colloid Interface Sci. 2019, 553, 530–539. [Google Scholar] [CrossRef] [PubMed]
  32. Vilé, G.; Di Liberto, G.; Tosoni, S.; Sivo, A.; Ruta, A.; Nachtegaal, M.; Clark, A.H.; Agnoli, S.; Zou, Y.; Savateev, A.; et al. Azide-alkyne click chemistry over a heterogeneous copper-based single-atom catalyst. ACS Catal. 2022, 12, 2947–2958. [Google Scholar] [CrossRef]
  33. Liu, J.; Zou, Y.; Cruz, D.; Savateev, A.; Antonietti, M.; G, Vilé. Ligand–metal charge transfer induced via adjustment of textural properties controls the performance of single-atom catalysts during photocatalytic degradation. ACS Appl. Mater. Interfaces 2021, 13, 25858–25867. [Google Scholar] [CrossRef] [PubMed]
  34. Yan, J.; Han, X.; Qian, J.; Liu, J.; Dong, X.; Xi, F. Preparation of 2D graphitic carbon nitride nanosheets by a green exfoliation approach and the enhanced photocatalytic performance. J. Mater. Sci. 2017, 52, 13091–13102. [Google Scholar] [CrossRef]
  35. Yang, Z.; Li, J.; Cheng, F.; Chen, Z.; Dong, X. BiOBr/protonated graphitic C3N4 heterojunctions: Intimate interfaces by electrostatic interaction and enhanced photocatalytic activity. J. Alloys Compd. 2015, 634, 215–222. [Google Scholar] [CrossRef]
  36. Zhao, Y.; Shi, R.; Bian, X.; Zhou, C.; Zhao, Y.; Zhang, S.; Wu, F.; Waterhouse, G.I.N.; Wu, L.Z.; Tung, C.H.; et al. Ammonia detection methods in photocatalytic and electrocatalytic experiments: How to improve the reliability of NH3 production rates? Adv. Sci. 2019, 6, 1802109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Shi, A.; Li, H.; Yin, S.; Hou, Z.; Rong, J.; Zhang, J.; Wang, Y. Photocatalytic NH3 versus H2 evolution over g-C3N4/CsxWO3: O2 and methanol tipping the scale. Appl. Catal. B Environ. 2018, 235, 197–206. [Google Scholar] [CrossRef]
  38. Yao, C.; Yuan, A.; Wang, Z.; Lei, H.; Zhang, L.; Guo, L.; Dong, X. Amphiphilic two-dimensional graphitic carbon nitride nanosheets for visible-light-driven phase-boundary photocatalysis. J. Mater. Chem. A 2019, 7, 13071–13079. [Google Scholar] [CrossRef]
  39. Tian, N.; Huang, H.; He, Y.; Guo, Y.; Zhang, T.; Zhang, Y. Mediator-free direct Z-scheme photocatalytic system: BiVO4/g-C3N4 organic–inorganic hybrid photocatalyst with highly efficient visible-light-induced photocatalytic activity. Dalton Trans. 2015, 44, 4297–4307. [Google Scholar] [CrossRef]
  40. Ma, T.Y.; Tang, Y.; Dai, S.; Qiao, S.Z. Proton-functionalized two-dimensional graphitic carbon nitride nanosheet: An excellent metal-/label-free biosensing platform. Small 2014, 10, 2382–2389. [Google Scholar] [CrossRef]
  41. Yan, H.; Yang, H. TiO2–g-C3N4 composite materials for photocatalytic H2 evolution under visible light irradiation. J. Alloys Compd. 2011, 509, L26–L29. [Google Scholar] [CrossRef]
  42. Zhu, B.; Xia, P.; Ho, W.; Yu, J. Isoelectric point and adsorption activity of porous g-C3N4. Appl. Surf. Sci. 2015, 344, 188–195. [Google Scholar] [CrossRef]
  43. Ye, L.; Liu, J.; Jiang, Z.; Peng, T.; Zan, L. Facets coupling of BiOBr-g-C3N4 composite photocatalyst for enhanced visible-light-driven photocatalytic activity. Appl. Catal. B Environ. 2013, 142–143, 1–7. [Google Scholar] [CrossRef]
  44. Wei, Y.; Cheng, G.; Xiong, J.; Zhu, J.; Gan, Y.; Zhang, M.; Li, Z.; Dou, S. Synergistic impact of cocatalysts and hole scavenger for promoted photocatalytic H2 evolution in mesoporous TiO2-NiSx hybrid. J. Energy Chem. 2019, 32, 45–56. [Google Scholar] [CrossRef]
  45. Kitano, M.; Kanbara, S.; Inoue, Y.; Kuganathan, N.; Sushko, P.V.; Yokoyama, T.; Hara, M.; Hosono, H. Electride support boosts nitrogen dissociation over ruthenium catalyst and shifts the bottleneck in ammonia synthesis. Nat. Commun. 2015, 6, 6731. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Pei, C.; Tan, J.; Li, Y.; Yao, G.; Jia, Y.; Ren, Z.; Liu, P.; Zhang, H. Effect of Sb-site nonstoichiometry on the structure and microwave dielectric properties of Li3Mg2Sb1−xO6 ceramics. J. Adv. Ceram. 2020, 9, 588–594. [Google Scholar] [CrossRef]
  47. Ruan, L.; Jia, Y.; Guan, J.; Xue, B.; Huang, S.; Wu, Z.; Li, G.; Cui, X. Highly piezocatalysis of metal-organic frameworks material ZIF-8 under vibration. Sep. Purif. Technol. 2022, 283, 120159. [Google Scholar] [CrossRef]
  48. Zhang, B.; Sun, R.; Wang, F.; Feng, T.; Zhang, P.; Luo, H. Pyroelectric properties of 91.5 Na0.5Bi0.5TiO3-8.5 K0.5Bi0.5TiO3 lead-free single crystal. J. Adv. Dielect. 2021, 11, 2150023. [Google Scholar] [CrossRef]
  49. 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]
  50. Dong, G.; Ho, W.; Wang, C. Selective photocatalytic N2 fixation dependent on g-C3N4 induced by nitrogen vacancies. J. Mater. Chem. A. 2015, 3, 23435–23441. [Google Scholar] [CrossRef]
  51. Liao, Y.; Lin, J.; Cui, B.; Xie, G.; Hu, S. Well-dispersed ultrasmall ruthenium on TiO2 (P25) for effective photocatalytic N2 fixation in ambient condition. J. Photochem. Photobiol. A Chem. 2020, 387, 112100. [Google Scholar] [CrossRef]
  52. Li, H.; Shang, J.; Shi, J.; Zhao, K.; Zhang, L. Facet-dependent solar ammonia synthesis of BiOCl nanosheets via a proton-assisted electron transfer pathway. Nanoscale 2016, 8, 1986–1993. [Google Scholar] [CrossRef] [PubMed]
  53. Sultana, S.; Mansingh, S.; Parida, K.M. Phosphide protected FeS2 anchored oxygen defect oriented CeO2NS based ternary hybrid for electrocatalytic and photocatalytic N2 reduction to NH3. J. Mater. Chem. A 2019, 7, 9145–9153. [Google Scholar] [CrossRef]
  54. Chen, L.; Wang, J.; Li, X.; Zhang, J.; Zhao, C.; Hu, X.; Lin, H.; Zhao, L.; Wu, Y.; He, Y. Facile preparation of Ag2S/KTa0.5Nb0.5O3 heterojunction for enhanced performance in catalytic nitrogen fixation via photocatalysis and piezo-photocatalysis. Green Energy Environ. 2022; in press. [Google Scholar] [CrossRef]
  55. Chen, L.; Zhang, W.; Wang, J.; Li, X.; Li, Y.; Hu, X.; Zhao, L.; Wu, Y.; He, Y. High piezo/photocatalytic efficiency of Ag/Bi5O7I nanocomposite using mechanical and solar energy for N2 fixation and methyl orange degradation. Green Energy Environ. 2021; in press. [Google Scholar] [CrossRef]
  56. 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]
  57. Jung, B.; Abu-Rub, F.; El-Ghenymy, A.; Deng, W.; Li, Y.; Batchelor, B.; Abdel-Wahab, A. Photocatalytic reduction of chlorate in aqueous TiO2 suspension with hole scavenger under simulated solar light. Emerg. Mater. 2021, 4, 435–446. [Google Scholar] [CrossRef]
  58. Diarmand-Khalilabad, H.; Habibi-Yangjeh, A.; Seifzadeh, D.; Asadzadeh-Khaneghah, S.; Vesali-Kermani, E. g-C3N4 nanosheets decorated with carbon dots and CdS nanoparticles: Novel nanocomposites with excellent nitrogen photofixation ability under simulated solar irradiation. Ceram. Int. 2019, 45, 2542–2555. [Google Scholar] [CrossRef]
  59. He, J.; Zhai, W.; Wang, S.; Wang, Y.; Li, W.; He, G.; Hou, X.; Liu, J.; He, Q. Persistently high Cr6+ removal rate of centi-sized iron turning owing to tribocatalysis. Mater. Today Phys. 2021, 19, 100408. [Google Scholar] [CrossRef]
Figure 1. SEM image of g-C3N4.
Figure 1. SEM image of g-C3N4.
Nanomaterials 12 01981 g001
Figure 2. XRD patterns of g-C3N4.
Figure 2. XRD patterns of g-C3N4.
Nanomaterials 12 01981 g002
Figure 3. FTIR spectra of g-C3N4.
Figure 3. FTIR spectra of g-C3N4.
Nanomaterials 12 01981 g003
Figure 4. XPS spectra of g-C3N4 sample: (a) survey, (b) C 1s, and (c) N 1s spectra.
Figure 4. XPS spectra of g-C3N4 sample: (a) survey, (b) C 1s, and (c) N 1s spectra.
Nanomaterials 12 01981 g004
Figure 5. Tribocatalytic N2 fixation performance of g-C3N4 with the different scavengers.
Figure 5. Tribocatalytic N2 fixation performance of g-C3N4 with the different scavengers.
Nanomaterials 12 01981 g005
Figure 6. Tribocatalytic N2 fixation performance of g-C3N4 with the different kinds of rods or without catalyst.
Figure 6. Tribocatalytic N2 fixation performance of g-C3N4 with the different kinds of rods or without catalyst.
Nanomaterials 12 01981 g006
Figure 7. Tribocatalytic N2 fixation performance of g-C3N4 under the different stirring speed.
Figure 7. Tribocatalytic N2 fixation performance of g-C3N4 under the different stirring speed.
Nanomaterials 12 01981 g007
Figure 8. Tribocatalytic N2 fixation performance of g-C3N4 with the different number of stirring rods.
Figure 8. Tribocatalytic N2 fixation performance of g-C3N4 with the different number of stirring rods.
Nanomaterials 12 01981 g008
Figure 9. The schematic diagram for the tribocatalytic mechanism of g-C3N4.
Figure 9. The schematic diagram for the tribocatalytic mechanism of g-C3N4.
Nanomaterials 12 01981 g009
Table 1. Summary of ammonia fixation performance of different catalysts and different catalytic methods.
Table 1. Summary of ammonia fixation performance of different catalysts and different catalytic methods.
CatalystsAmmonia Generation Rate/μmol·L−1·g−1·h−1Nitrogen SourceScavengerCatalytic Method
g-C3N4100.56airmethanolTribocatalysis [this work]
g-C3N4160airmethanolPhotocatalysis [50]
P2552N2waterPhotocatalysis [51]
BiOCl68.9N2methanolPhotocatalysis [52]
FeS2/CeO290N2waterPhotocatalysis [53]
KTa0.5Nb0.5O313.2airmethanolPiezocatalysis [54]
Ag/Bi5O7I65.4airwaterPiezocatalysis [55]
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.; Xu, T.; Ruan, L.; Guan, J.; Huang, S.; Dong, X.; Li, H.; Jia, Y. Strong Tribocatalytic Nitrogen Fixation of Graphite Carbon Nitride g-C3N4 through Harvesting Friction Energy. Nanomaterials 2022, 12, 1981. https://doi.org/10.3390/nano12121981

AMA Style

Wu Z, Xu T, Ruan L, Guan J, Huang S, Dong X, Li H, Jia Y. Strong Tribocatalytic Nitrogen Fixation of Graphite Carbon Nitride g-C3N4 through Harvesting Friction Energy. Nanomaterials. 2022; 12(12):1981. https://doi.org/10.3390/nano12121981

Chicago/Turabian Style

Wu, Zheng, Taosheng Xu, Lujie Ruan, Jingfei Guan, Shihua Huang, Xiaoping Dong, Huamei Li, and Yanmin Jia. 2022. "Strong Tribocatalytic Nitrogen Fixation of Graphite Carbon Nitride g-C3N4 through Harvesting Friction Energy" Nanomaterials 12, no. 12: 1981. https://doi.org/10.3390/nano12121981

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