Polyphenol-mediated interfacial deposition strategy for supported manganese oxide catalysts with excellent pollutant degradation performance

Dong Cheng Youyou Feng Bingxi Feng Ke Wang Guoxin Song Gen Wang Xiaoli Cheng Yonghui Deng Jing Wei

Citation:  Dong Cheng, Youyou Feng, Bingxi Feng, Ke Wang, Guoxin Song, Gen Wang, Xiaoli Cheng, Yonghui Deng, Jing Wei. Polyphenol-mediated interfacial deposition strategy for supported manganese oxide catalysts with excellent pollutant degradation performance[J]. Chinese Chemical Letters, 2024, 35(5): 108623. doi: 10.1016/j.cclet.2023.108623 shu

Polyphenol-mediated interfacial deposition strategy for supported manganese oxide catalysts with excellent pollutant degradation performance

English

  • Water contamination caused by pharmaceutical and personal care products (PPCPs) has become a serious environmental issue due to their great risk to ecosystem and public health [13]. Inefficient removal of PPCPS in municipal wastewater treatment plants (WWTPs) is one of the main reasons for the discharge of PPCPS [4,5]. As PPCPs cannot be effectively removed by conventional water treatment technologies like coagulation and sedimentation, more and more attention has been paid on advanced oxidation process (AOPs) for treating PPCPs [6,7]. In particular, catalytically activated peroxymonosulfate (PMS)-based AOPs has drawn increasing interest owing to the high oxidation capacity, mild operation condition and wide working pH range [8].

    Transition metal oxides (such as Co3O4, CuO, NiO and MnO2) has been extensively studied for PMS activation as they are highly active and natural abundance [912]. Among them, manganese oxides are more appealing due to the lower toxicity and rich chemical valence states [13,14]. Besides, metastable manganese intermediates can trigger a singlet oxygen (1O2) dominated non-radical oxidation process, which are able to selectively degrade target contaminants in complicated water matrices [15,16]. By now, various manganese oxides with different architectures, chemical states and crystal structures have been explored to active PMS for wastewater decontamination [17]. To maximize the catalytic efficiency, the manganese oxides were generally applied as a powdery catalyst during the treatment process [18,19]. When suspended in water, however, the powdery manganese oxides were readily to aggregate and lose, which greatly restricted their application in actual wastewater treatment [20]. Fixing metal oxides on monolithic substrates (such as minerals, metal foams and ceramic membranes) can effectively address the problem of catalyst loss or aggregation [2023]. Moreover, the fixed metal oxides could be installed in reactors of different shapes, which greatly favored their practical application [24]. Metal oxides were conventionally fixed through a slurry coating method by using a polymer binder, but suffered from the negative effect of polymer binder on the catalytic performance [25]. Recently, other strategies (such as chemical vapor deposition, hydrothermal method and spin coating) have also been explored to fix metal oxide catalysts but limited by the complicated synthesis process and extensive energy input [26]. Therefore, developing a facile and cost-effective strategy to fix manganese oxides on monolithic substrates is highly essential and desirable for sustainable water treatment.

    Recently, mussel-inspired chemistry has attracted intensive attentions due to its superior adhesive property on various substrates [2729]. Specially, Caruso's groups developed a simple, rapid and versatile coating strategy to deposit metal-polyphenol complex on various films and particles [3032]. This coating technique is powerful to deposit different kinds of metal species (e.g., Fe2+, Co2+, Ni2+, Mn2+) and polyphenol ligands on various substrates. Furthermore, polyphenols (e.g., tannic acid) not only show excellent adhesive property towards different interfaces, but also are one kind of reductant due to the abundant catechol groups in the molecular skeleton [3335]. For example, tannic acid (TA) has been widely used as an antioxidant. It has also been used as a reductant for the synthesis of noble metal nanoparticles [36,37]. Therefore, when combining the adhesive and reductive property of polyphenol, it is possible to deposit different kinds of metal oxides on various substrates. Till now, there are very few reports on the deposition of metal oxide on various interfaces based on the polyphenol redox chemistry.

    Herein, a polyphenol-mediated interfacial deposition strategy was developed to coat MnO2 on airstone, which is a kind of inert, chemically stable and low-cost minerals. The obtained composites were employed for PMS activation for sustainable degradation of PCPPs in the secondary effluent of WWTPs. The preparation process included the deposition of polyphenols (e.g., TA) on airstone substrates followed by interfacial redox reactions between TA molecules and KMnO4. Kinetic studies and mechanism investigations on PCPPs degradation over the fixed MnO2 were systematically studied using bis-phenol A (BPA) as a probe contaminant. Sustainable degradation of different kinds of PCPPs was explored via a flow-through reaction in a column reactor of the fixed MnO2, which revealed superior catalytic performance and long-term stability.

    TA is a typical plant polyphenol with abundant phenolic hydroxyl groups and possesses strong reductive capability [38,39]. Hence, TA molecules were used to react with oxidative KMnO4. MnO2 nanoparticles were obtained as indicated by the rapid color change of TA solution and scanning electron microscopy (SEM) image of the collected precipitate (Figs. S1 and S2 in Supporting information). During the synthesis, TA acted like a molecule glue and coated first on the surface of airstone substrates. Then the modified airstone was dispersed in KMnO4 solution, in which surface-coated TA molecules reduced KMnO4 to MnO2 nanoparticles to form MnO2 in-situ on the airstone. The color of the airstone substrates was turned from white to dark after the deposition process, revealing the successful fixing of MnO2 (Fig. 1a).

    Figure 1

    Figure 1.  (a) Schematic illustration for fixing MnO2 on airstone substrates. (b) SEM image and (c, d) magnified SEM images of the fixed MnO2. EDX elemental mapping of (e) Mn and (f) O, respectively. (g) HR-TEM image of the MnO2.

    SEM was applied to characterize the microstructure of the fixed MnO2. As shown in Fig. 1b, MnO2 nanoparticles with lose and porous structure were observed on the surface of airstone substrates whereas the pristine airstone displayed a smooth surface (Fig. S3 in Supporting information). The magnified SEM image further verified the porous structure (Figs. 1c and d). EDX mapping of the fixed MnO2 revealed that Mn and O elements were homogeneously dispersed (Figs. 1e and f). HR-TEM image of the MnO2 nanoparticles revealed low crystallinity (Fig. 1g). X-ray photoelectron spectroscopy (XPS) analysis and X-ray diffraction (XRD) pattern further confirmed low crystallinity of the fixed MnO2 (Fig. S4 in Supporting information). The loading of MnO2 on each gram of airstone substrate was determined to be ~0.07 mg according to the inductively coupled plasma (ICP) analysis.

    Depositing metal oxides on processable substrates is highly required in different applications as it can overcome the aggregation problem of nanosized metal oxides [40,41]. The surface-independent adhesion property of TA allows in-situ deposition of metal oxides on diverse substrates with differernt shapes. As shown in Fig. 2, successful deposition of MnO2 on different substrates was obtained, including metallic materials (e.g., stainless steel, iron sheet and copper foil), non-metallic materials (e.g., silicon wafer and glass flake) and polymer materials (e.g., rubber sheet and plastic piece). There was insignificant difference between the SEM images of pristine substates and TA-modified substrates (Fig. S5 in Supporting information). However, FT-IR spectrum of the TA-modified airstone displayed new peaks of TA at 3327, 1715, 1611 and 1317 cm−1, revealing the successful coating of TA (Fig. S6 in Supporting information). After reaction with KMnO4, the characteristic peaks of TA were vanished revealing there was no residual of TA in the fixed MnO2. The structure of deposited MnO2 is tunable by regulating the reaction parameters. Sheet-like MnO2 was formed at a low concentration of KMnO4 (< 1.0 mg/mL), while aggregated MnO2 particles was obtained when excess amount of KMnO4 (> 1.0 mg/mL) was applied (Figs. 2cf). N2 sorption isotherms of the MnO2 nanosheets exhibited type Ⅳ isotherms with a pore size centered at 5.1 nm (Fig. S7 in Supporting information), revealing a mesoporous structure. Aparting from MnO2, other oxide (such as Cr2O3) was also successfully fixed on the airstone as indicated by the SEM image and EDS mapping (Fig. S8 in Supporting information), revealing the versatility of the proposed interfacial deposition strategy.

    Figure 2

    Figure 2.  (a) Schematic illustration for in-situ deposition of MnO2 on different substrates. (b) Optical images of the substrates at different deposition stages: (1) iron sheet, (2) copper foil, (3) stainless steel sheet, (4) silicon wafer, (5) plastic piece, (6) rubber sheet and (7) glass flake. SEM images of MnO2 nanoparticles or nanosheets deposited on (c) iron sheet, (d) copper foil, (e) stainless steel sheet and (f) silicon wafer surface.

    Batch experiments were performed to investigate the oxidation capacity of the fixed MnO2/PMS system toward different organic contaminants. The removal of BPA in different reaction systems was compared firstly to study the degradation behavior of organic pollutant. As shown in Fig. 3a, only 8% of BPA was removed by PMS alone within 50 min, whereas complete removal of BPA was obtained in the fixed MnO2/PMS system. The adsorption of BPA on the fixed MnO2 was negligible and PMS catalyzed by airstone alone resulted in limited BPA removal (Fig. S9 in Supporting information). Besides, homogeneous catalysis of PMS by leached Mn2+ resulted in 20% removal of BPA. It indicated that the degradation of organic pollutant in the fixed MnO2/PMS system mainly occurred via heterogeneous reactions [42]. The degradation of BPA followed first-order kinetic model with a rate constant of 0.1 min−1 (R2 = 0.98) (Fig. 3a, inset), which was comparable with those of reported Mn-base metal oxides under similar reaction conditions (Table S1 in Supporting information). Interestingly, the degradation of BPA by 3.0 g of airstone with MnO2 (corresponding to ~0.2 mg of pristine MnO2) was much faster than that over 0.2 mg of powdery MnO2 (Fig. S10 in Supporting information). The enhanced PMS activation efficiency clearly manifested the advantage of the fixed MnO2, which relieved the aggregation problem of nanoparticles and thus afforded highly exposure of active sites. Besides, the structure of the fixed MnO2 provided open diffusion channels and efficient mass transport of PMS and organic pollutant, which also contributed to the enhanced removal of BPA [43].

    Figure 3

    Figure 3.  (a) Degradation of BPA in different reaction systems. (b) Degradation of different PPCPs in the fixed MnO2/PMS system. (c) TOC removal for different PPCPs. (d) Stability of the fixed MnO2 in consecutive runs. The influence of (e) Cl and NO3 and (f) humic acid on the degradation of BPA. (g) Degradation of BPA in the pH range from 3 to 11. (h) Continuous degradation of BPA in pure water and actual secondary effluent of WWPTs in a fixed-bed column of the fixed MnO2.

    Apart from the endocrine disrupting chemicals (e.g., BPA), the fixed MnO2/PMS system was also effective toward a wide spectrum of PCPPs including pharmaceuticals (e.g., acetaminophen, AAP), antibiotics (e.g., sulfamethoxazole, SMX) and organic dyes (e.g., methyl blue, MB). As shown in Fig. 3b, effective removal of AAP and MB was observed under the same reaction condition. Other organics like sulfamethoxazole (SMX), ibuprofen (IBU) and trichlorophenol (TCP) could also be degraded although a longer time was required (Fig. S11 in Supporting information). However, benzoic acid (BA) with electron-deficient groups could hardly be degraded under the same reaction condition, revealing the selective oxidation property of the fixed MnO2 + PMS system. The kinetic rate for AAP, SMX and MB was determined to be 0.09, 0.01 and 0.1 min−1, respectively (Fig. S12 in Supporting information). TOC removal of BPA, AAP, SMX, IBU, TCP and phenol was measured to be 77%, 64%, 54%, 28%, 52% and 57%, respectively (Fig. 3c), revealing PPCPs could be highly mineralized during the catalysis process. Besides, the fixed MnO2 exhibited superior stability and reusability in PMS activation. The removal efficiency of BPA was still as high as 92% after 5 cycle runs (Fig. 3d). The excellent stability should be partly ascribed to the low leaching of Mn2+ (Fig. S13 in Supporting information). On the other hand, the MnO2 was in-situ grown on airstone enabling strong adhesive of MnO2 nanoparticles on the substrate, which also contributed to the excellent stability. The stability of the fixed MnO2 was much higher than MnO2 powder (Fig. S14 in Supporting information). Additionally, the catalytic activity of the fixed MnO2 could be totally recovered via a facile calcination (Fig. S15 in Supporting information). The facile synthesis process, superior catalytic activity, excellent stability and simple regeneration strategy endowed the fixed MnO2 promising potential for PCPPS treatment. Fixing nanosized catalyst on processable substrates enables additional advantages of easy installation in reactors and promising potential for practical application. Herein, aiming to explore the application potential, the performance of the fixed MnO2 for treating PPCPs in the secondary effluent of WWTPs was investigated. Considering the complicated components of the secondary effluent and their possible quenching toward reactive oxygen species, the degradation of BPA in the presence of ubiquitous inorganic ions and natural organic matters (NOMs) was investigated first.

    As shown in Fig. 3e, the addition of Cl and NO3 hardly inhibited the degradation of BPA even when a high concentration of 50 mmol/L was applied. Other inorganic ions (e.g., SO42‒ and CO32‒) of high concentrations also exhibited negligible influence on BPA removal (Fig. S16 in Supporting information). The good resistance toward inorganic ions was ascribed to the 1O2-dominated nonradical oxidation mechanism (discussed later), which was moderate oxidative and less reactive with Cl, SO42‒, NO3 and CO32‒. Humic acid (HA) with a concentration from 5 mg/L to 50 mg/L showed insignificant influence on BPA removal as well (Fig. 3f), revealing the fixed MnO2/PMS system also restricted to NOMs. Besides, the fixed MnO2 had a wide working pH range (Fig. 3g). In comparison to the neutral condition, efficient removal of BPA was also obtained in acidic (pH 3, 5) and alkaline condition (pH 9, 11) with a slight decrease in kinetic rates. The fixed MnO2 was then employed to build a fixed-bed column reactor to treat PCPPs in the secondary effluent via a continuous flow-through reaction. The long-term stability of the column reactor was firstly studied by continuously degrading BPA in ultra-pure water. As shown in Fig. 3h, BPA was completely removed in the first one week with a high mineralization efficiency (TOC removal was ~70%), revealing a superior stability of the fixed MnO2. Then the ultra-pure water was replaced by secondary effluent of WWTP to explore the removal of BPA in actual wastewater. The removal efficiency of BPA was slightly decreased by ~3% probably due to the coverage of active sites by the background surroundings like NOMs [44]. Nevertheless, no further decline of BPA removal was observed and TOC was stably removed as well during one week of continuous reaction. Meanwhile, PMS was completely decomposed during the continuous run. The superior performance and excellent stability should be ascribed to the nonradical oxidation mechanism which was restricted to inorganic ions and NOMs. Moreover, the leaching of Mn2+ was low and met the requirement of the Environmental Quality Standards for Surface Water in China (EQSSW, GB3838–2002).

    Both radical (e.g., OH and SO4•‒) and nonradical oxidation (e.g., 1O2) have been proposed for organics degradation in PMS activation [13]. Herein, electron paramagnetic resonance (EPR) tests and chemical quenching experiments were performed to investigate the mechanism of PPCPs degradation. As shown in Fig. 4a, neither SO4•− nor OH was detected when 5, 5-dimethyl-1-pyrroline N-oxide (DMPO) was used as the spin trapping agent, indicating that no radicals were produced during the activation of PMS [14,17,45]. Chemical quenching experiments using methanol (MeOH) and tert-buty alcohol (TBA) showed that the degradation of BPA was not influenced even when the concentration of scavengers was as high as 1000 mmol/L. Besides, negligible degradation of benzoic acid (BA, a typical probe for radicals) was observed in the fixed MnO2/PMS system, further revealing the absence of radicals in organics degradation [46]. When 2, 2, 6, 6-tetramethyl-4-piperidone (TEMP) was used as the trapping agent, however, characteristic singles of TEMP-1O2 adducts was observed. Furthermore, the removal of BPA was strongly inhibited by introducing l-histidine (l-His, a typical chemical probe for 1O2). For example, the removal efficiency of BPA decreased by 80% when 50 mmol/L of l-His-was presented (Fig. 4b). It indicated that the degradation of PPCPs in the fixed MnO2/PMS system followed a 1O2-dominated nonradical oxidation pathway [47,48]. Chemical quenching experiments by benzoquinone (BQ, a scavenger for O2•−) revealed that the degradation of BPA was strongly inhibited after the introduction of BQ (Fig. S17 in Supporting information), indicating the 1O2 was originated from O2•− (2O2•− + H2O21O2 + H2O2 + 2OH) [22,23]. The variation of chemical states of Mn and O was analyzed to explore the mechanism of PMS activation over the fixed MnO2. After the catalysis reaction, the content of Mn4+ (643.2 eV) was decreased from 82.5% to 75.1% while the percentage of Mn3+ (641.1 eV) was increased from 17.5% to 24.9% (Fig. 4c). It indicated that the redox cycle of Mn4+/Mn3+ induced the activation of PMS activation [49,50]. XPS spectra of O 1s for used MnO2 showed that the percentage of surface oxygen was declined in comparison to that in the fresh MnO2 (Fig. 4d). In PMS activation, H2O molecules would be adsorbed on metal oxides surface and dissociated to form ≡Mn(Ⅳ)···OH and then participated in the subsequent metal oxide redox reaction [51]. In-situ attenuated total reflectance Fourier transform infrared spectroscopy (ATR-FTIR) was performed to further explore the activation process (Fig. 4e). For PMS alone, stretching vibration for S-O bond of SO42− and HSO5 was displayed at 1104 and 1256 cm−1, respectively. When contacted with the MnO2, the S-O bond of HSO5 was blue-shifted by 6 cm−1, suggesting the formation of chemical bond between MnO2 and HSO5 [22,52]. Moreover, the peak intensity of the S-O decreased gradually upon increasing the reaction time, indicating continuous decomposition of PMS molecules on the surface of MnO2. The results revealed that PMS was covalently bonded first with the MnO2, resulting in the formation of ≡Mn(Ⅳ)···O—O-SO3 complex on the surface (Eq. 1) (Fig. 4f). Then the complex reacted with another PMS molecule to produce 1O2 and participated into the degradation of micropollutants (Eqs. 2 and 3).

    (1)

    (2)

    (3)

    Figure 4

    Figure 4.  (a) Detection of SO4•−, OH and 1O2 via EPR tests. (b) Quenching effect of MeOH, TBA and l-His on the degradation of BPA. (c) XPS spectra for Mn 2p and (d) O 1s before and after the catalysis reaction. (e) ATR-FTIR spectra for MnO2/PMS at different times. (f) Schema for PMS activation and BPA degradation over the fixed MnO2.

    In summary, MnO2 was successfully fixed on airstone substrates via in-situ interfacial redox reactions between TA and KMnO4. Fixing MnO2 on the monolithic substrates allowed highly exposure of active sites and superior stability, and thus enabled superior catalytic performance toward PMS activation and PPCPs degradation with long-term stability. The degradation of PPCPs followed a 1O2-donminated nonradical oxidation pathway with strong anti-interference ability toward inorganic ions and natural organic matters. The fixed MnO2 could be installed in a column reactor for sustainable degradation of a wide spread of PPCPs in the secondary effluent of WWPTs. This work provided a reliable strategy for developing recoverable metal oxides with novel structures for sustainable water treatment.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    This work was financially supported by the National Key Research and Development Program (No. 2022YFE0100400), National Natural Science Foundation of China (No. 21701130), Key Basic Research Program of Science and Technology Commission of Shanghai Municipality (No. 20JC1415300), State Key Laboratory of Transducer Technology of China (No. SKT2207) and Key Research and Development Program of Shaanxi (No. 2021GY-225).

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2023.108623.


    1. [1]

      J.L. Liu, M.H. Wong, Environ. Int. 59 (2013) 208–224. doi: 10.1016/j.envint.2013.06.012

    2. [2]

      M.K. Shahid, A. Kashif, A. Fuwad, Y. Choi, Coord. Chem. Rev. 442 (2021) 213993. doi: 10.1016/j.ccr.2021.213993

    3. [3]

      M. Narayanan, M. El-Sheekh, Y. Ma, et al., Environ. Pollut. 300 (2022) 118922. doi: 10.1016/j.envpol.2022.118922

    4. [4]

      N.H. Tran, M. Reinhard, K.Y. Gin, Water Res. 133 (2018) 182–207. doi: 10.1016/j.watres.2017.12.029

    5. [5]

      H.B. Quesada, A.T.A. Baptista, L.F. Cusioli, et al., Chemosphere 222 (2019) 766–780. doi: 10.1016/j.chemosphere.2019.02.009

    6. [6]

      R.Y. Krishnan, S. Manikandan, R. Subbaiya, et al., Environ. Technol. Innov. 23 (2021) 101757. doi: 10.1016/j.eti.2021.101757

    7. [7]

      A. Kujawska, U. Kiełkowska, A. Atisha, et al., Sep. Purif. Technol. 290 (2022) 120797. doi: 10.1016/j.seppur.2022.120797

    8. [8]

      T. Liu, K. Cui, C.X. Li, et al., Chemosphere 311 (2023) 137084. doi: 10.1016/j.chemosphere.2022.137084

    9. [9]

      W. Li, X. He, B. Li, et al., Appl. Catal. B: Environ. 305 (2022) 121019. doi: 10.1016/j.apcatb.2021.121019

    10. [10]

      S. Wang, S. Gao, J. Tian, et al., J. Hazard. Mater. 387 (2020) 121995. doi: 10.1016/j.jhazmat.2019.121995

    11. [11]

      G. Wang, K. Wang, Z. Liu, et al., Appl. Catal. B: Environ. 325 (2023) 122359. doi: 10.1016/j.apcatb.2023.122359

    12. [12]

      R. Yang, Y. Fan, R. Ye, et al., Adv. Mater. 33 (2021) 2004862. doi: 10.1002/adma.202004862

    13. [13]

      M. Kohantorabi, G. Moussavi, S. Giannakis, Chem. Eng. J. 411 (2021) 127597.

    14. [14]

      Y. Liu, J. Luo, L. Tang, et al., ACS Catal. 10 (2020) 14857–14870. doi: 10.1021/acscatal.0c04049

    15. [15]

      Y. Yang, P. Zhang, K. Hu, et al., Appl. Catal. B: Environ. 315 (2022) 121593. doi: 10.1016/j.apcatb.2022.121593

    16. [16]

      L. Wang, J. Jiang, S.Y. Pang, et al., Chem. Eng. J. 352 (2018) 1004–2013. doi: 10.3390/ma11061004

    17. [17]

      Y. Yang, P. Zhang, K. Hu, et al., Appl. Catal. B: Environ. 286 (2021) 119903. doi: 10.1016/j.apcatb.2021.119903

    18. [18]

      F. Pan, H. Ji, P. Du, et al., J. Hazard. Mater. 402 (2021) 123779. doi: 10.1016/j.jhazmat.2020.123779

    19. [19]

      Y. Liu, R. Qu, X. Li, et al., J. Colloid Interface Sci. 579 (2020) 412–424. doi: 10.1016/j.jcis.2020.06.073

    20. [20]

      F. Ghanbari, M. Moradi, Chem. Eng. J. 310 (2017) 41–62. doi: 10.1016/j.cej.2016.10.064

    21. [21]

      X. Dong, B. Ren, X. Zhang, et al., Appl. Catal. B: Environ. 272 (2020) 118971. doi: 10.1016/j.apcatb.2020.118971

    22. [22]

      X. Dong, X. Duan, Z. Sun, et al., Appl. Catal. B: Environ. 261 (2020) 118214. doi: 10.1016/j.apcatb.2019.118214

    23. [23]

      G. Wang, Y. Zhang, L. Ge, et al., J. Hazard. Mater. 429 (2022) 128282. doi: 10.1016/j.jhazmat.2022.128282

    24. [24]

      G. Wang, L. Ge, Z. Liu, et al., Chem. Eng. J. 431 (2022) 134026. doi: 10.1016/j.cej.2021.134026

    25. [25]

      D. Chen, Q. Wang, R. Wang, et al., J. Mater. Chem. A 3 (2015) 10158–10173. doi: 10.1039/C4TA06923D

    26. [26]

      S. Mathur, T. Ruegamer, N. Donia, J. Nanosci. Nanotechnol. 8 (2008) 2597–2603. doi: 10.1166/jnn.2008.633

    27. [27]

      Y. Liu, K. Ai, L. Lu, Chem. Rev. 114 (2014) 5057–5115. doi: 10.1021/cr400407a

    28. [28]

      J. Saiz-Poseu, J. Mancebo-Aracil, F. Nador, et al., Angew. Chem. Int. Ed. 58 (2019) 696–714. doi: 10.1002/anie.201801063

    29. [29]

      E. Filippidi, T.R. Cristiani, C.D. Eisenbach, et al., Science 358 (2017) 502–505. doi: 10.1126/science.aao0350

    30. [30]

      H. Ejima, J.J. Richardson, K. Liang, et al., Science 341 (2013) 154–157. doi: 10.1126/science.1237265

    31. [31]

      H. Geng, Q.Z. Zhong, J. Li, et al., Chem. Rev. 122 (2022) 11432–11473. doi: 10.1021/acs.chemrev.1c01042

    32. [32]

      J. Zhou, Z. Lin, M. Penna, et al., Nat. Commun. 11 (2020) 4804. doi: 10.1038/s41467-020-18589-0

    33. [33]

      J. Guo, Y. Ping, H. Ejima, et al., Angew. Chem. Int. Ed. 53 (2014) 5546–5551. doi: 10.1002/anie.201311136

    34. [34]

      J. Qin, G. Liang, B. Feng, et al., Chin. Chem. Lett. 32 (2021) 842–848. doi: 10.1016/j.cclet.2020.05.021

    35. [35]

      S. Quideau, D. Deffieux, C. Douat-Casassus, et al., Angew. Chem. Int. Ed. 50 (2011) 586–621. doi: 10.1002/anie.201000044

    36. [36]

      Y. Feng, P. Li, J. Wei, Coord. Chem. Rev. 468 (2022) 214649. doi: 10.1016/j.ccr.2022.214649

    37. [37]

      B. Feng, Y. Feng, Y. Li, et al., ACS Sens. 7 (2022) 3963–3972. doi: 10.1021/acssensors.2c02232

    38. [38]

      J. Wei, G. Wang, F. Chen, et al., Angew. Chem. Int. Ed. 57 (2018) 9838–9843. doi: 10.1002/anie.201805781

    39. [39]

      G. Wang, J. Qin, X. Zhou, et al., Adv. Funct. Mater. 28 (2018) 1806144. doi: 10.1002/adfm.201806144

    40. [40]

      X. Dong, Z. Chen, A. Tang, et al., Adv. Funct. Mater. 32 (2022) 2111565. doi: 10.1002/adfm.202111565

    41. [41]

      L. Zhu, J. Ji, J. Liu, et al., Angew. Chem. Int. Ed. 59 (2020) 13968–13976. doi: 10.1002/anie.202006059

    42. [42]

      C. Huang, Y. Wang, M. Gong, et al., Sep. Purif. Technol. 230 (2020) 115877. doi: 10.1016/j.seppur.2019.115877

    43. [43]

      J. Huang, Y. Dai, K. Singewald, et al., Chem. Eng. J. 370 (2019) 906–915. doi: 10.1016/j.cej.2019.03.238

    44. [44]

      Q.Z. Zhong, S. Li, J. Chen, et al., Angew. Chem. Int. Ed. 58 (2019) 12563–12568. doi: 10.1002/anie.201907666

    45. [45]

      J. Ji, Q. Yan, P. Yin, et al., Angew. Chem. Int. Ed. 60 (2021) 2903–2908. doi: 10.1002/anie.202013015

    46. [46]

      S. Zhu, X. Li, J. Kang, et al., Environ. Sci. Technol. 53 (2019) 307–315. doi: 10.1021/acs.est.8b04669

    47. [47]

      X. Li, J. Wang, X. Duan, et al., ACS Catal. 11 (2021) 4848–4861. doi: 10.1021/acscatal.0c05089

    48. [48]

      P. Shao, J. Tian, F. Yang, et al., Adv. Funct. Mater. 28 (2018) 1705295. doi: 10.1002/adfm.201705295

    49. [49]

      G.X. Huang, C.Y. Wang, C.W. Yang, et al., Environ. Sci. Technol. 51 (2017) 12611–12618. doi: 10.1021/acs.est.7b03007

    50. [50]

      J. Huang, S. Zhong, Y. Dai, et al., Environ. Sci. Technol. 52 (2018) 11309– 11318. doi: 10.1021/acs.est.8b03383

    51. [51]

      Z. Chen, S. Bi, G. Zhao, et al., Sci. Total Environ. 711 (2020) 134715. doi: 10.1016/j.scitotenv.2019.134715

    52. [52]

      S. Yang, S. Xu, J. Tong, et al., Appl. Catal. B: Environ. 295 (2021) 120291. doi: 10.1016/j.apcatb.2021.120291

  • Figure 1  (a) Schematic illustration for fixing MnO2 on airstone substrates. (b) SEM image and (c, d) magnified SEM images of the fixed MnO2. EDX elemental mapping of (e) Mn and (f) O, respectively. (g) HR-TEM image of the MnO2.

    Figure 2  (a) Schematic illustration for in-situ deposition of MnO2 on different substrates. (b) Optical images of the substrates at different deposition stages: (1) iron sheet, (2) copper foil, (3) stainless steel sheet, (4) silicon wafer, (5) plastic piece, (6) rubber sheet and (7) glass flake. SEM images of MnO2 nanoparticles or nanosheets deposited on (c) iron sheet, (d) copper foil, (e) stainless steel sheet and (f) silicon wafer surface.

    Figure 3  (a) Degradation of BPA in different reaction systems. (b) Degradation of different PPCPs in the fixed MnO2/PMS system. (c) TOC removal for different PPCPs. (d) Stability of the fixed MnO2 in consecutive runs. The influence of (e) Cl and NO3 and (f) humic acid on the degradation of BPA. (g) Degradation of BPA in the pH range from 3 to 11. (h) Continuous degradation of BPA in pure water and actual secondary effluent of WWPTs in a fixed-bed column of the fixed MnO2.

    Figure 4  (a) Detection of SO4•−, OH and 1O2 via EPR tests. (b) Quenching effect of MeOH, TBA and l-His on the degradation of BPA. (c) XPS spectra for Mn 2p and (d) O 1s before and after the catalysis reaction. (e) ATR-FTIR spectra for MnO2/PMS at different times. (f) Schema for PMS activation and BPA degradation over the fixed MnO2.

  • 加载中
计量
  • PDF下载量:  2
  • 文章访问数:  473
  • HTML全文浏览量:  2
文章相关
  • 发布日期:  2024-05-15
  • 收稿日期:  2023-03-25
  • 接受日期:  2023-05-29
  • 修回日期:  2023-05-22
  • 网络出版日期:  2023-06-03
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章