Next Article in Journal
Biodesulfurization of Dibenzothiophene by Decorating Rhodococcus erythropolis IGTS8 Using Montmorillonite/Graphitic Carbon Nitride
Next Article in Special Issue
Catalytic Ozonation of Norfloxacin Using Co-Mn/CeO2 as a Multi-Component Composite Catalyst
Previous Article in Journal
CO2 Electroreduction on Carbon-Based Electrodes Functionalized with Molecular Organometallic Complexes—A Mini Review
Previous Article in Special Issue
Kinetics and Mechanisms of Cr(VI) Removal by nZVI: Influencing Parameters and Modification
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

N Doped Activated Biochar from Pyrolyzing Wood Powder for Prompt BPA Removal via Peroxymonosulfate Activation

1
College of Materials Science and Engineering, Nanjing Forestry University, Nanjing 210037, China
2
Jiangsu Co-Innovation Center of Efficient Processing and Utilization of Forest Resources, International Innovation Center for Forest Chemicals and Materials, College of Materials Science and Engineering, Nanjing Forestry University, Nanjing 210037, China
*
Author to whom correspondence should be addressed.
Submission received: 28 October 2022 / Revised: 11 November 2022 / Accepted: 14 November 2022 / Published: 16 November 2022

Abstract

:
In the present study, nitrogen doped biochar (N-PPB) and nitrogen doped activated biochar (AN-PPB) were prepared and used for removing bisphenol A (BPA) in water through activating peroxymonosulfate. It was found from the results that N-PPB exhibited superior catalytic performance over pristine biochar since nitrogen could brought about abundant active sites to the biochar structure. The non-radical singlet oxygen (1O2) was determined to be the dominant active species responsible for BPA degradation. Having non-radical pathway in the N-PPB/PMS system, the BPA degradation was barely influenced by many external environmental factors including solution pH value, temperature, foreign organic, and inorganic matters. Furthermore, AN-PPB had richer porosity than N-PPB, which showed even faster BPA removal efficiency than N-PPB through an adsorptive/catalytic synergy. The finding of this study introduces a novel way of designing hieratical structured biochar catalysts for effective organic pollutant removal in water.

1. Introduction

In recent years, advanced oxidation processes (AOPs) based on persulfate activations have been widely studied for catalytic degradation of organic contaminants [1,2]. Amongst various approaches for activating persulfate, transitional metal ions or metal compounds exhibit superior efficiency [3]. Nevertheless, it is intractable to overcome the metal ion leaching issue which can cause serious environmental pollution problems [4]. Recent studies have shown that carbonaceous materials can also effectively activate persulfate [5]. Thereinto, wood-derived biochar produced from pyrolyzing biomass precursor in low-oxygen or oxygen-free atmosphere has received tremendous research interest due to the advantages of abundant existence, low cost, low biotoxicity, and good stability, as well as numerous active oxygen-containing functional groups [6,7].
However, pristine biochar has low the degradation efficiency and poor recycling performance due to its limited active sites within skeletons for persulfate activation [8]. Nitrogen doping is considered to be an effective method to improve the catalytic capability of biochar through introducing diverse N-containing functional groups (pyridine N, pyrrole N, graphitic N) on the surface of biochar [9], which can not only change the electron density of local carbon atoms and increase electron mobility, but also increase edge defects of biochar [10]. At the same time, N atoms can also adjust the surface chemical properties through changing the charge/spin distribution of biochar, improve the adsorption capacity of persulfate, and add extra active sites [11], which are all in favor of catalytic degradations of organic pollutants [12]. On the other hand, biochar can be converted to activated carbon by an activation agent, through which biochar can have more surface oxygen functional groups, higher defects degree of internal carbon structure, and richer porosity [13]. Furthermore, the resulted activated biochar can be endowed with immensely increase the specific surface area along with hierarchical pore structures, which promote the adsorption capability of the biochar [14]. Therefore, it is potential to obtain N-doped activated biochar which have high organic pollutant removal efficiency in water through adsorption/degradation synergy through combining N doping and activating modification treatments to wood-based biochar.
Thus, in this study, N-doped biochar catalysts with different pyrolysis temperatures (450 °C, 600 °C, 750 °C, 900 °C) were prepared using industrial poplar powder as precursor and urea as nitrogen source. After comparing the catalytic performance through degrading bisphenol A (BPA) via activating peroxymonosulfate (PMS), the biochar pyrolyzed at 750 °C with optimal performance was then selected to prepare N-doped activated biochar using KOH as activating agent. The performance of N-doped activated biochar/PMS system towards BPA degradation was studied systematically afterwards. Furthermore, the recyclability and actual water adaptability of the reaction system were also investigated. The BPA degradation mechanisms in the reaction system were also analyzed.

2. Results and Discussion

The surface morphology structure difference between pristine poplar powder and N-PPB investigated in terms of SEM with the results shown in Figure 1. It could be observed from Figure 1a that pristine poplar powder displayed a smooth strip shape. After pyrolysis treatment, and nitrogen doping, N-PPB750 exhibited a rougher structure with richer porosity (Figure 1b), which was considerably distinguished from pristine poplar powder. At the same time, the particles of N-PPB750 became smaller compared with that of pristine poplar powder, which was attributed to the decomposition of lignocellulose component in the poplar powder during pyrolysis process [15]. The pyrolysis treatment also promoted the exfoliation of the carbon layers, which lead to the formation of abundant pores. The EDS elemental mapping diagrams shown in Figure 1d–f demonstrated that several main elements including C, N, and O were uniformly distributed N-PPB750. The contents of three major elements shown in Table 1 illustrated that a quite high content of N element (17.98 wt%) imported by urea precursor was anchored in the biochar structures.
The structures of N-PPB samples were then identified through XRD analysis with the results shown in Figure 2a. It could be seen that the pristine poplar powder had three wide diffraction peaks at 2θ = 15°, 22°, and 35°, which were affiliated to the Cellulose I structure of typical wood based biomass materials [16]. After pyrolysis, these three peaks were replaced by two broad peaks located at ~26° and 42° in N-PPB samples, which corresponded to (002) and (110) planes of classic carbon materials with graphitic structures [17]. These results further indicated the formation of fine graphitic structure in the prepared N-PPB samples. Additionally, many sharp small peaks in the range of 22–40° were also observed in all tested samples, which were related to the mineral salts such as calcium carbonate and silica existed in natural poplar powder [18]. These peaks became weaker with the elevation of pyrolysis temperature, which might be because some of the mineral salts were evaporated at higher temperature.
Figure 2b–f shows the Raman spectra of the N-PPB samples. As shown in Figure 2b, two characteristic peaks located at 1360 cm−1 and 1580 cm−1 could be observed for all the N doped biochar samples, which were ascribed to the characteristic D band and G band of a graphitic carbonaceous material [19]. This meant poplar wood could be converted to carbonaceous biochar when the pyrolysis temperature was higher than 450 °C. After the Raman spectra was deconvoluted, it was seen that the spectrum of each N-PPB sample were consisted of seven characteristic peaks which were ascribed to were divided into the SR band representing C-H on aromatic rings (1086 cm−1), the S band representing Caromatic-Calkyl (1170 cm−1), the SL band representing Caryl-O-Calkyl (1245 cm−1), the D band representing the defective and disordered arrangement of carbon atoms (1320 cm−1), the VR band representing to amorphous carbon and deterioration of crystallinity (1390 cm−1), the VL band representing the semicircle ring breathing (1468 cm−1), and the G band representing in-plane stretching vibration of C atomic sp2 hybridization (1576 cm−1), respectively [20]. The area ratio between D band and G band (AD/AG) was generally used to characterize the defect degree of the graphitic structure in a carbonaceous material [21]. It was seen that the AD/AG value increased from 0.83 to 1.15 with the increase in the pyrolysis temperature, indicating that higher pyrolysis temperature brought about more defect to the biochar. In the meantime, the area ratio between D band and (VR+VL) bands (AD/AV) represented the defect density of boundary edges in the carbon. It was seen that all the biochar samples had an AD/AV value of lower than 3.5, indicating that the prepared N-PPB samples were considered to contain a higher density defect of boundary edges with more unsaturated carbon atoms and a lower degree of graphitization [22].
XPS analysis was then conducted to further explore the surface element composition of the prepared N-PPB. As illustrated in Figure 3a, the main elements of C, N, and O were all observed in the spectra of the biochar samples, in which the concentration of N decreased significantly with the increase in the pyrolysis temperature. Figure 3b–d reveals the deconvoluted C1s, O1s, and N1s spectra. It was observed from Figure 3b that C1s spectrum was mainly composed of four peaks located at 284.5 eV, 285.2 eV, 286.2 eV, and 288.8 eV, which represented the C=C/C-C, C–O, C=O and COOH bonding, respectively [23,24]. Similarly, the O1s spectrum could be divided into C=O (530.9 eV), C-O (532.2 eV), and COOH (533.6 eV) (Figure 3c), and N1s spectra could be deconvoluted into four peaks corresponding to pyridinic—N (398.3 eV), pyrrolic—N (399.5 eV), graphite—N (400.4 eV), and −NOx (401.4 eV) (Figure 3d) [25]. The detailed concentrations of respective bonding were listed in Table 2. As shown, the concentration of C=C/C-C decreased with the increment of pyrolysis temperature from 450 °C to 900 °C, which was in agreement with the Raman results. Correspondingly, the proportion of C=O bonding increased with the increase in the pyrolysis temperature, which could be resulted from the conversion of C=C/C-C bonding. As for N1s bonding, the conversion of pyrrolic N or pyridine N to graphitic N could be obviously seen.
The catalytic performance of the prepared N-PPB samples was investigated through degrading BPA in water with the results shown Figure 4a. As illustrated, the PPB samples showed negligible adsorption activity towards BPA since the biochar was not activated. Moreover, nitrogen doping could significantly enhance the PMS activation capability of the biochar, and all the N-PPB samples (0.5 g/L) could completely degrade 0.02 mM BPA in the solution. Within the same time span, PPB without N doping could only degrade ~10% of BPA. This indicated that N doping significantly enriched the active sites on PPB surface, which accelerated the PMS activation efficiency of the biochar sample. It was also seen that the activation capability of the N-PPB sample increased with the pyrolysis temperature, and when pyrolysis temperature increased to 600 °C or higher, the catalytic performance of the corresponding PPB did not increase much. The TOC results obtained from N-PPB600/PMS system indicated that when the reaction time was extended to 30 min, ~58% of the TOC could be removed. This result implied that the BPA could be degraded into small molecules by the prepared N-PPB catalysts and finally mineralized into inorganic molecules.
N-PPB600 was selected for the following test. The impact of reaction parameters, including BPA concentration, catalyst dosage, and PMS concentration on the performance of N-PPB600, was then investigated with the results shown in Figure 4b–d. It was revealed that the catalytic performance of N-PPB600 had a positive correlation with the catalyst dosage, and had a negative correlation with the BPA concentration. It was easy to understand that increased catalyst dosage could increase the number of active sites for PMS activation, which led to the increase in active species for BPA degradation [26]. Conversely, when the concentrations of both N-PPB600 and PMS were fixed, a constant number of active species was insufficient to degrade increased amount of BPA molecules, which resulted in the performance decline of N-PPB600. While when the concentration of PMS was increased in the reaction solution, the performance of N-PPB600 increased when PMS concentration increased from 1.5 mM to 2.0 mM, and declined when further increasing the PMS concentration. This might because an excessive amount of PMS would also consume the generated active species, which inhibited the BPA degradation [27].
The impact of solution parameters on the performance of N-PPB600 was investigated afterwards. Figure 5a shows the influence of solution pH on the BPA degradation efficiency. Overall, the catalytic performance of N-PPB600 was not much influenced by the fluctuation of solution pH value, indicating promising performance stability of the catalyst. It was further observed from Figure 5b that the activity of N-PPB600 boosted with the increase in the system temperature, indicating the endothermic nature of the PMS activation process [28].
Figure 6 shows the impact of foreign matters including inorganic anions, natural organic matters and complex water matrices on the performance of the biochar catalyst. As illustrated in Figure 6a–d, all inorganic anions could influence the activity of N-PPB600. Specifically, H2PO4, HCO3, and NO3 inhibited the BPA degradation efficiency since these anions would occupy the catalyst surface and hinder the contact between the active sites in the catalyst and PMS, resulting lower active species generation rate [29]. Unlike the other three anions, the dosage of Cl sped up the BPA degradation rate, and the increased amount of Cl would accelerate the BPA degradation rate more intensely. This was because Cl could react with PMS and active species and generate more chlorine based species [30], which, as a result, boosted the removal efficiency of BPA in the reaction system. It was also seen from Figure 6e–f that both humic acid and natural water had an impediment effect on the catalytic activity of N-PPB600, which was ascribed to the similar reason to the anions as discussed above. It has to be noted that even the concentration of the anions and humic acid increased to as high as 50 mM; 0.02 mM of BPA could still be effectively degraded with high removal rate, implying promising water matrix adaptability of the prepared catalyst.
The active species generated in the reaction system which dominated the BPA degradation was investigated through scavenging test with the results shown in Figure 7a. It was observed that both methanol (MeOH) and tert-butyl alcohol (TBA) did not show any inhibition effect to BPA degradation, indicating that no radical species was generated in the reaction system. When the trapping agent was changed to L-histidine and para-benzoquinone (p-BQ), the BPA degradation could be completely inhibited. Since these two agents were both non-radical scavenger which could effectively trap singlet oxygen (1O2) in the solution [31], this meant the N-PPB600/PMS system was a 1O2 dominated system for BPA degradation. This could also explain why N-PPB600 could kept its catalytic activity at various water matrices since 1O2 was not much influenced by the variation of external environment.
The ESR results shown in Figure 7b also illustrated that when using TMP as the trapping agent, a clear triplet peak which was ascribed to the typical TMP-1O2 adduct could be observed, indicating that abundant 1O2 was existed in the N-PPB600/PMS system. Many previous studies indicated that the electron-rich Lewis basic sites in biochar could transfer the electrons to PMS and caused the generation of non-radical 1O2 [32]. Since nitrogen doping could introduce abundant electron-rich nitrogen containing groups to the biochar including pyridinic N and pyrrolic N [33], the N-PPB600 could exhibited higher PMS activating efficiency than the pristine biochar.
The catalytic stability of N-PPB600 was investigated afterwards. It was seen from Figure 7c that the catalytic capability of the biochar gradually reduced after four consecutive runs, which indicated that the active sites on the catalyst surface was continuously consumed when activating PMS. Thus, the XPS spectra of the recycled N-PPB600 was investigated to illustrate the PMS activation mechanism with the results shown in Figure 8. As shown in Figure 8a, the N element was significantly consumed after consecutive BPA degradation cycles, indicating that N doping played a critical role of improving the catalytic properties of the biochar. From the deconvoluted C1s, N1s, and O1s spectra shown in Figure 8b–d, it was seen that the concentrations of C=O, pyridinic N and graphitic N were reduced after the reaction cycles, implying that these bonding participated the PMS activation and were consumed gradually, which were in accordance with many previous studies [34].
To improve the BPA degradation efficiency of the biochar, AN-PPB600 was prepared to enhance the adsorption capability of the biochar. It was seen that AN-PPB600 had much higher specific surface area and pore volume than N-PPB600 after KOH activation (Table 3), which indicated that AN-PPB600 should have better adsorption capability than N-PPB600. As illustrated in Figure 9, ~80% of BPA could be adsorbed by AN-PPB600 at the first 30 min, while no more than 10% of BPA was adsorbed by N-PP600. When PMS was dosed, 0.5 g/L of AN-PPB600 could completely remove BPA with high concentration of 0.1 mM through adsorptive/catalytic synergy within 3 min. Conversely, N-PPB600 could only degrade ~15% of BPA within the same time span. This indicated that through designing hieratical structured biochar with both high adsorptive and catalytic capability, organic pollutants in water could be promptly removed through the synergetic process.

3. Experimental

3.1. Reagents and Chemicals

Poplar powder (300 mesh) was obtained from Yixing wood powder factory (Linyi, China). PMS (Oxone, KHSO5·0.5KHSO4·0.5K2SO4) and BPA (99%) were purchased from Sigma-Aldrich (St. Louis, MO, USA). Urea (CH4N2O, 99%) was provided from Nanjing Chemical Reagent Co., Ltd. (Nanjing, China). 2,2,6,6-tetramethyl-4-piperidine (TMP, 98%) was acquired from Aladdin (Shanghai, China). All the other chemicals and reagents were of analytical grade and used without further purification. Double distilled water (H2O) was utilized throughout the whole experiments.

3.2. Preparation of the Catalysts

Typically, 3.0 g urea was first dissolved into 20 mL H2O, after which 1.0 g poplar powder was added. The mixture was magnetically stirred 70 °C until all the H2O was evaporated. Afterwards, the obtained powder was put into a tubular furnace, heated to desired temperature (450 °C, 600 °C, 750 °C and 900 °C) under N2 atmosphere at 5 °C /min, pyrolyzed at respective temperature for 3 h. After cooling to room temperature, the obtained solid was further washed with H2O for several times until the pH value of H2O reached ~7. After being dried in an oven at 60 °C for 24 h, N-doped biochar derived from poplar powder was obtained, and was designated to be N-PPBx, where x represented the pyrolysis temperature. For comparison, activated N-doped biochar was synthesized via a similar route except that 2.0 g KOH was also dissolved in urea solution in the first step, which was named as AN-PPBx.

3.3. Characterizations

The morphology of biochar was investigated by scanning electron microscope (SEM, JEOL SEM 6490, Tokyo, Japan). The crystalline structures were studied by X-ray diffraction diffractometer (XRD, Rigaku Smartlab, Tokyo, Japan). Raman spectra of the prepared samples was investigated by a Raman spectrometer equipped with an Ar laser (532 nm, 180 mW) (Horiba Jobin Yvon HR800Tokyo, Japan). X-ray photoelectron spectroscopy (XPS, Kratos Ultra DLD, Warwick, UK) was used to detect surface chemistry and element composition. Electron spin resonance spectroscopy (ESR, Bruker EMX-10/12, Bremen, Germany) was used to identify generated reactive oxygen species (ROS) in the biochar/PMS system. Electrochemical measurements were carried out with a CHI760E electrochemical workstation. The concentration of BPA was analyzed by high performance liquid chromatography (HPLC, Dionex Ultimate 3000, Sunnyvale, CA USA) equipped with C18 column (5 μm particle size, 250 × 4.6 mm) and UV detector at 278 nm. Total organic carbon (TOC) was measured by a MutiN/C 2100 analyzer (Jena, Germany).

3.4. Degradation Process

All the catalytic degradation experiments were conducted in a glass beaker containing 100 mL BPA solution with continuous mechanical stirring. In brief, a certain amount of biochar sample was first dispersed in H2O and sonicated for 30 min to obtain a uniform dispersion. Afterwards, BPA solution with pre-determined concentration was obtained through injecting a certain amount of BPA stock solution into the dispersion. Before PMS was dosed to initiate BPA degradation, the mixture was stirred for 30 min to achieve the adsorption/desorption equilibrium. During the BPA degradation process, 1 mL of solution sample was periodically extracted from the reaction system, filtered with a 0.22 μm polytetrafluoroethylene membrane, and quenched with 0.5 mL methanol. The initial pH value of the solution was measured after PMS and catalyst were added, and was adjusted by NaOH (1 M) and H2SO4 (1 M). In a recycling test, the used biochar was separated from the reaction system through centrifuging, which was then rinsed 3 times with clean H2O and dried at 60 °C for 12 h before being dosed to initiate next BPA degradation cycle. The ROS produced during oxidative degradation of pollutants were identified by scavenging experiments using methanol (MeOH), isobutanol (TBA), p-benzoquinone (p-BQ), and L-histdine (His) as scavengers. All experiments were conducted in triplicates, and the data were the mean values with standard deviations.

4. Conclusions

To conclude, N-PPB catalysts were prepared and used for the removal of BPA in water through activating PMS. The results shows that nitrogen doping could significantly enhance the catalytic performance of pristine biochar, in which 0.02 mM of BPA could be completely degraded within 5 min by 0.5 g/L of biochar catalyst. The scavenging test showed that non-radical 1O2 was the main active species which dominated the BPA degradation. Due to this reason, the N-PPB/PMS system showed promising water matrix adaptability. To enhance the BPA removal efficiency, KOH was introduced to activate N-PPB, and the results illustrated that AN-PPB could promptly remove high concentration BPA in water through adsorptive/catalytic synergy. This study provides a new thinking of preparing biochar catalyst with heteroatom doing and hieratical structure design, which could remove organic pollutants in water with high efficiency.

Author Contributions

Investigation, writing—original draft preparation, data curation, H.L.; visualization, formal analysis, G.X.; writing—review and editing, supervision, funding acquisition, L.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of Jiangsu Province, China (BK20201385).

Data Availability Statement

Data are available upon request.

Acknowledgments

The Advanced Analysis and Testing Center of Nanjing Forestry University is acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shi, J.; Dai, B.; Fang, X.; Xu, L.; Wu, Y.; Lu, H.; Cui, J.; Han, S.; Gan, L. Waste preserved wood derived biochar catalyst for promoted peroxymonosulfate activation towards bisphenol A degradation with low metal ion release: The insight into the mechanisms. Sci. Total Environ. 2022, 813, 152673. [Google Scholar] [CrossRef] [PubMed]
  2. Lv, Y.; Zong, L.; Liu, Z.; Du, J.; Wang, F.; Zhang, Y.; Ling, C.; Liu, F. Sequential separation of Cu(II)/Ni(II)/Fe(II) from strong-acidic pickling wastewater with a two-stage process based on a bi-pyridine chelating resin. Chin. Chem. Lett. 2021, 32, 2792–2796. [Google Scholar] [CrossRef]
  3. Sethi, Y.A.; Panmand, R.P.; Ambalkar, A.; Kulkarni, A.K.; Patil, D.R.; Gunjal, A.R.; Gosavi, S.W.; Kulkarni, M.V.; Kale, B.B. In situ preparation of CdS decorated ZnWO4 nanorods as a photocatalyst for direct conversion of sunlight into fuel and RhB degradation. Sustain. Energy Fuels 2019, 3, 793–800. [Google Scholar] [CrossRef]
  4. Pan, Y.; Bu, Z.; Sang, C.; Guo, H.; Zhou, M.; Zhang, Y.; Tian, Y.; Cai, J.; Wang, W. EDTA enhanced pre-magnetized Fe0/H2O2 process for removing sulfamethazine at neutral pH. Sep. Purif. Technol. 2020, 250, 117281. [Google Scholar] [CrossRef]
  5. Lu, H.; Gan, L. Catalytic Degradation of Bisphenol A in Water by Poplar Wood Powder Waste Derived Biochar via Peroxymonosulfate Activation. Catalysts 2022, 12, 1164. [Google Scholar] [CrossRef]
  6. Zhao, Z.; Zhang, X.; Lin, Q.; Zhu, N.; Gui, C.; Yong, Q. Development and investigation of a two-component adhesive composed of soybean flour and sugar solution for plywood manufacturing. Wood Mater. Sci. Eng. 2022, 1–9. [Google Scholar] [CrossRef]
  7. Zuo, S.; Zhang, W.; Wang, Y.; Xia, H. Low-Cost Preparation of High-Surface-Area Nitrogen-Containing Activated Carbons from Biomass-Based Chars by Ammonia Activation. Ind. Eng. Chem. Res. 2020, 59, 7527–7537. [Google Scholar] [CrossRef]
  8. Bhatia, S.K.; Gurav, R.; Choi, T.-R.; Kim, H.J.; Yang, S.-Y.; Song, H.-S.; Park, J.Y.; Park, Y.-L.; Han, Y.-H.; Choi, Y.-K.; et al. Conversion of waste cooking oil into biodiesel using heterogenous catalyst derived from cork biochar. Bioresour. Technol. 2020, 302, 122872. [Google Scholar] [CrossRef]
  9. You, Y.; Zhao, Z.; Song, Y.; Li, J.; Li, J.; Cheng, X. Synthesis of magnetized nitrogen-doped biochar and its high efficiency for elimination of ciprofloxacin hydrochloride by activation of peroxymonosulfate. Sep. Purif. Technol. 2020, 258, 117977. [Google Scholar] [CrossRef]
  10. Baştürk, E.; Alver, A. Modeling azo dye removal by sono-fenton processes using response surface methodology and artificial neural network approaches. J. Environ. Manag. 2019, 248, 109300. [Google Scholar] [CrossRef]
  11. Ye, S.; Zeng, G.; Tan, X.; Wu, H.; Liang, J.; Song, B.; Tang, N.; Zhang, P.; Yang, Y.; Chen, Q.; et al. Nitrogen-doped biochar fiber with graphitization from Boehmeria nivea for promoted peroxymonosulfate activation and non-radical degradation pathways with enhancing electron transfer. Appl. Catal. B Environ. 2020, 269, 118850. [Google Scholar] [CrossRef]
  12. Huang, B.-C.; Jiang, J.; Huang, G.-X.; Yu, H.-Q. Sludge biochar-based catalysts for improved pollutant degradation by activating peroxymonosulfate. J. Mater. Chem. A 2018, 6, 8978–8985. [Google Scholar] [CrossRef]
  13. Liu, X.; Zuo, S.; Cui, N.; Wang, S. Investigation of ammonia/steam activation for the scalable production of high-surface area nitrogen-containing activated carbons. Carbon 2022, 191, 581–592. [Google Scholar] [CrossRef]
  14. Gao, H.; Zuo, S.; Wang, S.; Xu, F.; Yang, M.; Hu, X. Graphitic crystallite nanomaterials enable the simple and ultrafast synthesis of resorcinol-formaldehyde carbon aerogel monoliths. Carbon 2022, 194, 220–229. [Google Scholar] [CrossRef]
  15. Gan, L.; Zhong, Q.; Geng, A.; Wang, L.; Song, C.; Han, S.; Cui, J.; Xu, L. Cellulose derived carbon nanofiber: A promising biochar support to enhance the catalytic performance of CoFe2O4 in activating peroxymonosulfate for recycled dimethyl phthalate degradation. Sci. Total Environ. 2019, 694, 133705. [Google Scholar] [CrossRef]
  16. Geng, A.; Meng, L.; Han, J.; Zhong, Q.; Li, M.; Han, S.; Mei, C.; Xu, L.; Tan, L.; Gan, L. Highly efficient visible-light photocatalyst based on cellulose derived carbon nanofiber/BiOBr composites. Cellulose 2018, 25, 4133–4144. [Google Scholar] [CrossRef]
  17. Yu, Y.; Liu, Y.; Wu, X.; Weng, Z.; Hou, Y.; Wu, L. Enhanced visible light photocatalytic degradation of metoprolol by Ag–Bi2WO6–graphene composite. Sep. Purif. Technol. 2015, 142, 1–7. [Google Scholar] [CrossRef]
  18. Jia, H.; Zhao, S.; Zhu, K.; Huang, D.; Wu, L.; Guo, X. Activate persulfate for catalytic degradation of adsorbed anthracene on coking residues: Role of persistent free radicals. Chem. Eng. J. 2018, 351, 631–640. [Google Scholar] [CrossRef]
  19. Hu, W.; Tong, W.; Li, Y.; Xie, Y.; Chen, Y.; Wen, Z.; Feng, S.; Wang, X.; Li, P.; Wang, Y.; et al. Hydrothermal route-enabled synthesis of sludge-derived carbon with oxygen functional groups for bisphenol A degradation through activation of peroxymonosulfate. J. Hazard. Mater. 2019, 388, 121801. [Google Scholar] [CrossRef]
  20. Cheng, X.; Guo, H.; Zhang, Y.; Wu, X.; Liu, Y. Non-photochemical production of singlet oxygen via activation of persulfate by carbon nanotubes. Water Res. 2017, 113, 80–88. [Google Scholar] [CrossRef]
  21. Hou, P.; Xing, G.; Tian, L.; Zhang, G.; Wang, H.; Yu, C.; Li, Y.; Wu, Z. Hollow carbon spheres/graphene hybrid aerogels as high-performance adsorbents for organic pollution. Sep. Purif. Technol. 2018, 213, 524–532. [Google Scholar] [CrossRef]
  22. Liang, J.; Xu, X.; Zaman, W.Q.; Hu, X.; Zhao, L.; Qiu, H.; Cao, X. Different mechanisms between biochar and activated carbon for the persulfate catalytic degradation of sulfamethoxazole: Roles of radicals in solution or solid phase. Chem. Eng. J. 2019, 375, 121908. [Google Scholar] [CrossRef]
  23. Pereira, A.T.; Henriques, P.C.; Costa, P.C.; Martins, M.C.L.; Magalhães, F.D.; Gonçalves, I.C. Graphene oxide-reinforced poly(2-hydroxyethyl methacrylate) hydrogels with extreme stiffness and high-strength. Compos. Sci. Technol. 2019, 184, 107819. [Google Scholar] [CrossRef]
  24. Pan, Y.; Wang, Q.; Zhou, M.; Cai, J.; Tian, Y.; Zhang, Y. Kinetic and mechanism study of UV/pre-magnetized-Fe0/oxalate for removing sulfamethazine. J. Hazard. Mater. 2020, 398, 122931. [Google Scholar] [CrossRef] [PubMed]
  25. He, J.; Zhou, Q.; Chen, S.; Tian, M.; Zhang, C.; Sun, W. Interfacial microstructures and adsorption mechanisms of benzohydroxamic acid on Pb2+-activated cassiterite (110) surface. Appl. Surf. Sci. 2020, 541, 148506. [Google Scholar] [CrossRef]
  26. Liu, C.; Liu, L.; Tian, X.; Wang, Y.; Li, R.; Zhang, Y.; Song, Z.; Xu, B.; Chu, W.; Qi, F.; et al. Coupling metal–organic frameworks and g-C3N4 to derive Fe@N-doped graphene-like carbon for peroxymonosulfate activation: Upgrading framework stability and performance. Appl. Catal. B-Environ. 2019, 255, 117763. [Google Scholar] [CrossRef]
  27. Wang, J.; Wang, S. Preparation, modification and environmental application of biochar: A review. J. Clean. Prod. 2019, 227, 1002–1022. [Google Scholar] [CrossRef]
  28. Tao, W.; Duan, W.; Liu, C.; Zhu, D.; Si, X.; Zhu, R.; Oleszczuk, P.; Pan, B. Formation of persistent free radicals in biochar derived from rice straw based on a detailed analysis of pyrolysis kinetics. Sci. Total Environ. 2020, 715, 136575. [Google Scholar] [CrossRef]
  29. Wu, J.; Grant, P.; Li, X.; Noble, A.; Aggarwal, V.K. Catalyst-Free Deaminative Functionalizations of Primary Amines by Photoinduced Single-Electron Transfer. Angew. Chem. Int. Ed. 2019, 58, 5697–5701. [Google Scholar] [CrossRef] [Green Version]
  30. Duan, X.; Sun, H.; Wang, S. Metal-Free Carbocatalysis in Advanced Oxidation Reactions. Acc. Chem. Res. 2018, 51, 678–687. [Google Scholar] [CrossRef]
  31. Liu, J.; Huang, S.; Chen, K.; Wang, T.; Mei, M.; Li, J. Preparation of biochar from food waste digestate: Pyrolysis behavior and product properties. Bioresour. Technol. 2020, 302, 122841. [Google Scholar] [CrossRef]
  32. Han, R.; Fang, Y.; Sun, P.; Xie, K.; Zhai, Z.; Liu, H.; Liu, H. N-Doped Biochar as a New Metal-Free Activator of Peroxymonosulfate for Singlet Oxygen-Dominated Catalytic Degradation of Acid Orange 7. Nanomaterials 2021, 11, 2288. [Google Scholar] [CrossRef]
  33. Hu, Y.; Chen, D.; Zhang, R.; Ding, Y.; Ren, Z.; Fu, M.; Cao, X.; Zeng, G. Singlet oxygen-dominated activation of peroxymonosulfate by passion fruit shell derived biochar for catalytic degradation of tetracycline through a non-radical oxidation pathway. J. Hazard. Mater. 2021, 419, 126495. [Google Scholar] [CrossRef]
  34. Lata, H.; Garg, V.; Gupta, R. Adsorptive removal of basic dye by chemically activated Parthenium biomass: Equilibrium and kinetic modeling. Desalination 2008, 219, 250–261. [Google Scholar] [CrossRef]
Figure 1. SEM images of (a) PPB-750, (b) N-PPB750, (c) selected area of N-PPB750 (red box) for EDS elemental mapping and EDS elemental mapping of (d) C, (e) O, (f) N.
Figure 1. SEM images of (a) PPB-750, (b) N-PPB750, (c) selected area of N-PPB750 (red box) for EDS elemental mapping and EDS elemental mapping of (d) C, (e) O, (f) N.
Catalysts 12 01449 g001
Figure 2. (a) XRD patterns, (b) Raman spectra, and (cf) deconvoluted Raman spectra of N-PPB450, N-PPB600, N-PPB750, N-PPB900.
Figure 2. (a) XRD patterns, (b) Raman spectra, and (cf) deconvoluted Raman spectra of N-PPB450, N-PPB600, N-PPB750, N-PPB900.
Catalysts 12 01449 g002
Figure 3. (a) Wide-scan XPS spectra, high resolution (b) C 1s, (c) N 1s, (d) O 1s spectra of N-PPB catalysts.
Figure 3. (a) Wide-scan XPS spectra, high resolution (b) C 1s, (c) N 1s, (d) O 1s spectra of N-PPB catalysts.
Catalysts 12 01449 g003
Figure 4. (a) Catalytic performance of N-PPB samples ([catalyst]0 = 0.5 g/L, [BPA]0 = 0.02 mM, [PMS]0 = 1.5 mM), impact of (b) catalyst dosage ([BPA]0 = 0.02 mM, [PMS]0 = 1.5 mM), (c) BPA concentration ([catalyst]0 = 0.5 g/L, [PMS]0 = 1.5 mM), and (d) PMS concentration on the performance of N-PPB600 ([catalyst]0 = 0.5 g/L, [BPA]0 = 0.02 mM).
Figure 4. (a) Catalytic performance of N-PPB samples ([catalyst]0 = 0.5 g/L, [BPA]0 = 0.02 mM, [PMS]0 = 1.5 mM), impact of (b) catalyst dosage ([BPA]0 = 0.02 mM, [PMS]0 = 1.5 mM), (c) BPA concentration ([catalyst]0 = 0.5 g/L, [PMS]0 = 1.5 mM), and (d) PMS concentration on the performance of N-PPB600 ([catalyst]0 = 0.5 g/L, [BPA]0 = 0.02 mM).
Catalysts 12 01449 g004
Figure 5. Impact of (a) pH value and (b) solution temperature on the performance of N-PPB600 ([catalyst]0 = 0.5 g/L, [BPA]0 = 0.02 mM, [PMS]0 = 1.5 mM).
Figure 5. Impact of (a) pH value and (b) solution temperature on the performance of N-PPB600 ([catalyst]0 = 0.5 g/L, [BPA]0 = 0.02 mM, [PMS]0 = 1.5 mM).
Catalysts 12 01449 g005
Figure 6. Impact of (a) H2PO4, (b) Cl, (c) HCO3, (d) NO3, (e) humic acid, and (f) natural water on the performance of N-PPB600 ([catalyst]0 = 0.5 g/L, [BPA]0 = 0.02 mM, [PMS]0 = 1.5 mM).
Figure 6. Impact of (a) H2PO4, (b) Cl, (c) HCO3, (d) NO3, (e) humic acid, and (f) natural water on the performance of N-PPB600 ([catalyst]0 = 0.5 g/L, [BPA]0 = 0.02 mM, [PMS]0 = 1.5 mM).
Catalysts 12 01449 g006
Figure 7. (a) BPA degradation in N-PPB600/PMS system at the existence of various scavengers, (b) ESR spectra test, (c) recyclability test for N-PPB600 ([catalyst]0 = 0.5 g/L, [BPA]0 = 0.02 mM, [PMS]0 = 1.5 mM).
Figure 7. (a) BPA degradation in N-PPB600/PMS system at the existence of various scavengers, (b) ESR spectra test, (c) recyclability test for N-PPB600 ([catalyst]0 = 0.5 g/L, [BPA]0 = 0.02 mM, [PMS]0 = 1.5 mM).
Catalysts 12 01449 g007
Figure 8. (a) Wide-scan XPS spectra, (b) C 1s, (c) N 1s, (d) O 1s spectra of used N-PPB600.
Figure 8. (a) Wide-scan XPS spectra, (b) C 1s, (c) N 1s, (d) O 1s spectra of used N-PPB600.
Catalysts 12 01449 g008
Figure 9. BPA degradation in N-PPB600/PMS and AN-PPB600/PMS systems ([catalyst]0 = 0.5 g/L, [BPA]0 = 0.02 mM, [PMS]0 = 1.5 mM).
Figure 9. BPA degradation in N-PPB600/PMS and AN-PPB600/PMS systems ([catalyst]0 = 0.5 g/L, [BPA]0 = 0.02 mM, [PMS]0 = 1.5 mM).
Catalysts 12 01449 g009
Table 1. Elemental ratio of N-PPB750 (C, O, N).
Table 1. Elemental ratio of N-PPB750 (C, O, N).
ElementWeight%Atomic%
C K50.5256.40
N K17.9817.21
O K31.5026.39
Totals100.00
Table 2. Bonding ratio of N-PPB samples based on deconvoluted C1s and N1s XPS peak.
Table 2. Bonding ratio of N-PPB samples based on deconvoluted C1s and N1s XPS peak.
C1sN1s, %
C=C/C-CC-OC=O–COOHPyridinic NPyroolic NGraphitic N–NOx
N-PPB45033.2%32.0%23.7%11.1%47.9%22.8%20.2%9.1%
N-PPB60030.7%34.2%26.8%8.3%46.6%17.2%21.4%14.8%
N-PPB75027.2%34.5%28.1%10.2%45.1%16.8%22.6%15.5%
N-PPB90031.7%33.4%26.4%8.5%31.0%17.6%27.5%23.9%
Table 3. Specific surface area and pore volume of N-PPB600 and AN-PPB600.
Table 3. Specific surface area and pore volume of N-PPB600 and AN-PPB600.
SampleSpecific Surface Area (m2·g−1)Pore Volume (cm3 g−1)
N-PPB6001410.203
AN-PPB6008551.683
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Lu, H.; Xu, G.; Gan, L. N Doped Activated Biochar from Pyrolyzing Wood Powder for Prompt BPA Removal via Peroxymonosulfate Activation. Catalysts 2022, 12, 1449. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12111449

AMA Style

Lu H, Xu G, Gan L. N Doped Activated Biochar from Pyrolyzing Wood Powder for Prompt BPA Removal via Peroxymonosulfate Activation. Catalysts. 2022; 12(11):1449. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12111449

Chicago/Turabian Style

Lu, Haiqin, Guilu Xu, and Lu Gan. 2022. "N Doped Activated Biochar from Pyrolyzing Wood Powder for Prompt BPA Removal via Peroxymonosulfate Activation" Catalysts 12, no. 11: 1449. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12111449

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