Facile preparation of α-MnO2 nanowires for assembling free-standing membrane with efficient Fenton-like catalytic activity

Yufei Zhen Zhiqiang Sun Ziye Jia Caihong Liu Shishu Zhu Xueyan Li Wei Wang Jun Ma

Citation:  Yufei Zhen, Zhiqiang Sun, Ziye Jia, Caihong Liu, Shishu Zhu, Xueyan Li, Wei Wang, Jun Ma. Facile preparation of α-MnO2 nanowires for assembling free-standing membrane with efficient Fenton-like catalytic activity[J]. Chinese Chemical Letters, 2023, 34(3): 107664. doi: 10.1016/j.cclet.2022.07.007 shu

Facile preparation of α-MnO2 nanowires for assembling free-standing membrane with efficient Fenton-like catalytic activity

English

  • Fenton-like advanced oxidation processes (AOPs) have been receiving lots of attentions in the fields of water treatment due to their low cost and high oxidation potentials [1-4]. Metal oxides exhibited high reactivities towards the catalyzed H2O2 propagations, which initiated radical chain reactions for pollutant degradation [5]. Manganese dioxide (MnO2) was an efficient heterogenous catalyst to activate such processes, activating H2O2 to produce large amounts of superoxide anions (O2•−) and hydroxyl radicals (OH) [6, 7]. Moreover, this catalyzed reaction would be promoted when framing MnO2 to a nanosized shape with different crystal forms and facet exposures [8]. However, it was not practical to recycle and separate the suspended MnO2 nanomaterials, thus limiting large-scale applications [9]. Some techniques of solving this issue have been developed, such as magnetic modification [10], support loading [11], and membrane construction [12]. Among these, assembling nanomaterials into macroscopic membrane was considered as the most promising engineered technology. The catalytic membrane not only promoted the potential of nanostructured MnO2 separation and recovery, but also enhanced the Fenton-like reaction activity by confining such process in a nanodomain membrane pore [13, 14]. The simultaneous filtration and reaction in confined space overcame the mass transfer limitations which often encountered in bulk diffusive mode operations [15, 16]. Therefore, exploring facile approaches to prepare catalytic membrane using MnO2 nanomaterials was urgent for practical applications.

    Nanomaterials, enabling to be assembled into membrane, should be designed with special architectures to guarantee mechanical strength during the membrane filtration [17]. Assembling one dimensional (1D) MnO2 nanowire into two-dimensional (2D) membrane by vacuum filtration was extensively used due to its advantages including easy-operation and liable controllability [18]. In that case, it was prerequisite to prepare the well-structured MnO2 nanowires. Some previous studies employed hydrothermal methods to accelerate the redox reactions of Mn hydrated ions by adding oxidative reagents (e.g., HNO3 or (NH4)2S2O8) to obtain MnO2 nanowires [12, 19, 20]. However, for these current approaches, the reactions always concluded various extra components (mixed crystal) and byproducts (e.g., MnOOH) [21]. Such issues would not only affect the purity of obtained MnO2 nanowires, but also raised the difficulty of reaction control. In addition, since the vacuum filtration method required very high aspect ratio and good dispersion of MnO2 nanowires in the solvents, few approaches could be applied in preparing such featured MnO2 nanowires [12]. Therefore, studies on facile synthetic methods to obtain MnO2 nanowires for catalytic membrane assembling and practical applications of water treatment were in great demands.

    Herein, we developed a facile method to prepare ultralong α-MnO2 nanowires with high purity and aspect ratio under 140 ℃ for 12 h through hydrothermal reaction. Specially, the C2H5OH and CH3COOK were used as reductive and control reagents respectively to react with KMnO4 as manganese source. Compared to previous studies [12, 19, 20], this strategy exhibited three advantages as follows: (1) This synthesis reaction was mild, easily controlled, and insensitive to temperature, reagents ratio, and reaction time; (2) the byproducts in reaction system were less and safer than those in other methods; (3) all reaction reagents were cheap, readily available, and environmentally friendly. We further investigated the formation mechanism of α-MnO2 nanowires via various characterizations. Moreover, we assembled these α-MnO2 nanowires into free-standing membrane as a Fenton catalyst for organic degradation in water. By the comparison with α-MnO2 nanowire powder in aqueous solution, the possible mechanism of enhanced pollution removals and the application potential of α-MnO2 nanowire membrane in Fenton-like oxidation were comprehensively explored.

    Details of experimental procedures and characterizations were exhibited in Texts S1–S4 (Supporting information). The phase purity and crystal structure of the prepared MnO2 were examined by XRD in Fig. 1a. All the diffraction peaks could be exclusively indexed as tetragonal α-MnO2 (JCPDS No. 44-0141, space group I4/m, a = b = 9.784 Å, c = 2.863 Å), which was a type of the well-defined 2 × 2 tunnel structure with a tunnel size of 4.6 Å and was composed of edge shared MnO6 octahedra with corner sharing double chains. No other characteristic peaks of the impurities were detected in the spectrum, indicating high phase purity of the obtained α-MnO2 [20]. The elemental composition of α-MnO2 was detected by EDS mapping (Fig. 1b). The atomic ratio of Mn to O was very close to 2, confirming that the highly purified α-MnO2 was successfully synthesized. The morphology of α-MnO2 was characterized by SEM and TEM. As shown in Fig. 1c, the α-MnO2 were homogenously overlapped in a 1D nanowire shape as expected. The nanowire morphology grew into as long as tens of micrometers with diameter of ~50 nm (Fig. S1 in Supporting information). Such ultralong nanowires were ensured to form a stable membrane with high porosity and flexibility by easily interconnecting with each other [22]. Notably, a little number of bundles made up of ultralong nanowires (inset in Fig. 1c) parallelly coexisted with together, which indicated some details of "nucleation-growth" mechanism [21]. The ultralong nanowires were typical mesoporous structure (type II isotherm, Fig. S2 in Supporting information) and the BET surface area was up to 34.2 m2/g, which would endow the as-prepared α-MnO2 with excellent catalytic performance [23]. With the help of HR-TEM, the deeper microstructure of α-MnO2 nanowires was investigated (Fig. 1d). The clear crystal lattices signified that the ultralong nanowire was single-crystalline with crystal orientation along [001] direction (c axis). Along the growth axis, the interplanar spacing was 0.69 nm, corresponding to the (110) facets. Two other interplanar spacings between the adjacent lattice planes were also measured to be 0.31 and 0.23 nm, which could be assigned to the (310) and (211) facets of α-MnO2. Moreover, the SAED pattern displayed orderly spot arrays which were indexed to the (110), (200), (310) and (211) planes. Comparing the intensities of XRD diffraction peaks, it could be found that the intensities corresponding to exposed facets in α-MnO2 were relatively stronger. As known, the exposed facets had high percentage of unsaturated atoms and great potential to react with low percentage of unsaturated atoms [24, 25]. In addition, the high-index exposed facets were usually composed of high density of under-coordinated atoms, such as steps, edges, and kinks, serving as active sites. The successful synthesis of α-MnO2 with high-index exposed facets would contribute to enhancing the catalytic activity [20].

    Figure 1

    Figure 1.  (a) XRD pattern, (b) EDS spectrum, (c) SEM images and (d) HR-TEM images and SAED pattern of prepared MnO2 nanowires.

    The formation mechanism of α-MnO2 nanowires was systematically discussed and proposed. C2H5OH and CH3COOK were identified as the key factors in the formation of ultralong nanowire shaped α-MnO2. C2H5OH acted as a mild reducing agent to react with manganese source, MnO4. Redox potential of MnO4/MnO2 (0.588 V in neutral condition) was only slightly higher than that of CH3COOH/C2H5OH (0.31 V in neutral condition) which acted as the reductive electrode couple (Eq. 1) [26]. Such approximately standard electrode potentials made the reaction mild and slow, facilitating the "nucleation-growth" process [12, 27]. However, the reactive solution containing C2H5OH and MnO4 could only form flower shaped α-MnO2 instead of nanowire shaped (Fig. 2a). The lateral growth of α-MnO2 needed to be restricted to form nanowires. As seen in Eq. 1, the reaction between C2H5OH and MnO4 did not generate steric reagent, which led to lateral growth into the flower shape of obtained MnO2.

    (1)

    (2)

    (3)

    (4)

    Figure 2

    Figure 2.  SEM images of MnO2 nanowires prepared under different conditions: (a) CH3CH2OH as reductive reagent at 140 ℃ for 12 h, (b) CH3CHO as reductive reagent at 140 ℃ for 12 h, (c) CH3CH2OH as reductive reagent and H2SO4 as control reagent at 140 ℃ for 12 h, (d) CH3CH2OH as reductive reagent and CH3COOH as control reagent at 140 ℃ for 12 h.

    Previous study suggested that CH3COOH could hinder the growth of lateral nanostructure through steric effects due to the formation of hydrogen bonds with -OH groups on MnO2 [28]. For example, when employing CH3CHO as a substitute of C2H5OH to reduce MnO4, the formed MnO2 transferred from flower-shape to nanowire-shape (Fig. 2b) due to the formation of byproduct CH3COOH (Eq. 2). Moreover, both adding H+ (i.e., H2SO4) which enhanced the hydrolysis of generated CH3COO to be CH3COOH and directly adding CH3COOH in C2H5OH/MnO4 system (Eq. 3) could further lead to the nanowire-shaped growth of MnO2 (Figs. 2c and d). However, in the conditions outlined above, the obtained MnO2 nanowires were short and seriously bonded with each other (Figs. 2bd). This was because the gap of redox potential between electrode couple (e.g., CH3COOH/CH3CHO (−0.12 V) < C2H5OH/CH3COOH (0.31 V) in the neutral condition, MnO4/MnO2 (0.588 V in neutral condition) < MnO4/MnO2 (1.510 V in acidic condition)) in the reactive system became larger under these conditions, resulting in rapid formation of MnO2 nuclei [29]. Too many nuclei of MnO2 lowered the development of ultralong dendritic structure and further restricted the nanowire-shaped growth of MnO2. To create a mild redox condition, CH3COOK, which was also the only byproduct in the reaction system, was introduced into the reactive solution and the ultralong α-MnO2 nanowires were obtained (Fig. 1c and Fig. S1). On the one hand, the hydrolysis of CH3COO could generate CH3COOH at the beginning of this reaction (Eq. 4). On the other hand, such process could produce extra OH which further decreased the gap of redox potential between CH3COOH/CH3CH2OH and MnO4/MnO2 (0.564 V in base condition). This method created both steric configuration and mild reactive condition for well control of the ultralong growth of α-MnO2 nanowires, and was insensitive to temperature, reagent ratio, and reaction time (Fig. S3 in Supporting information).

    The formation mechanism of ultralong α-MnO2 nanowires using our method was depicted in Fig. 3. Firstly, the primary nucleation of MnO2 occurred slowly in the solution through the reaction between MnO4 and C2H5OH in weak base condition and formed three-dimension MnO2 cores under hydrothermal condition. Then the hydrolysis of CH3COOK leaded to the production of CH3COOH attached to the surface of MnO2 nuclei. When such reaction proceeded in a mild manner under steric effect of CH3COOH, the MnO2 intermediates grew along the c axis and formed ultralong nanowires slowly and homogenously. This approach was not only environmentally friendly and easily implemented, but also forward-looking for developing other ultralong 1D nanostructures of metal oxides.

    Figure 3

    Figure 3.  Schematic illustration for the formation mechanism of α-MnO2 nanowires.

    MnO2 can enable Fenton-like chemistry and activate hydrogen peroxide (H2O2) to produce hydroxyl radicals (OH) for organic pollutant degradation [5]. The catalytic performance of catalysts is intensively related to their application forms [30]. Herein, the catalytic performance of the relevant free-standing α-MnO2 membrane (Figs. 4a and b) assembled by ultralong α-MnO2 nanowires towards BPA degradation was deeply investigated. To exhibit the advantage of membrane catalyzation with continuous flow through MnO2 membrane (MnO2-M, Fig. S4 in Supporting information) over traditional heterogeneous catalytic process with bulk MnO2 particle suspension (MnO2-P, Fig. S4), systemically comparison of the removal process was conducted between MnO2-M and MnO2-P towards BPA. As shown in Fig. 4c, the removal of BPA by MnO2-M, MnO2-P, or H2O2 solution alone within 60 min was negligible (less than 10%) due to respective limited reactivity. When MnO2 was coupled with H2O2, BPA removal was enhanced significantly because Fenton-like reaction was initiated and a large amount of OH were generated to oxidize the aqueous organic contaminants. Particularly, the reactivity of MnO2-M (kobs = 0.1112 min−1) confined within nanowire membrane was obviously higher than that of MnO2-P (kobs = 0.0441 min−1) in bulk solution (Fig. 4c). Moreover, the catalytic reactivity of MnO2-M was considerable for most of the model phenolic pollutants (Fig. 4d), in addition of the limited Mn2+ release less than 0.05 mg/L in reactive solution (Fig. S6 in Supporting information), indicating a preferable potential in practical application.

    Figure 4

    Figure 4.  (a, b) Digital photo and SEM images of obtained MnO2 membrane, (c) BPA removal efficiency and (d) reactive rate constant for organic pollutants of MnO2 membrane catalyzed H2O2 system. Condition: solution volume, 50 mL; MnO2-M/MnO2-P, 1 mg; H2O2, 50 mg/L; BPA, 2 mg/L; pH, 6.75; for continuous flow reaction (MnO2-M), recycle solution speed, 50 mL/min; for batch reaction, stirring speed, 300 r/min.

    Due to the numerous channels or pores in membrane structure, the nanoconfinement effect is often hypothesized as a crucial factor for tuning membrane performance [30]. Therefore, it is speculated that the nanoconfinement effect could interpret this excellent performance when numerous channels (Fig. 4b) were observed to be developed within the ultralong α-MnO2 nanowire membrane. Previous studies demonstrated that when chemical reactions occurred within a small or nanoscale confined space, their activity would be improved by altering the dielectric constant of water molecules, enrichment and exitance of reactants, or energy barriers [13, 14, 29, 31]. Based on the second-order rate of BPA with OH (k2 of 6.9 × 109 L mol−1 s−1) [32], the steady-state OH concentration in MnO2-M/H2O2 system (0.267 pmol/L) was approximately 2.5-fold higher than that in MnO2-P/H2O2 system (0.107 pmol/L) [33]. This result indicated that more OH were exposed within the MnO2-M to attack organic substrates due to the occurring nanoconfinement effect. When the overlapped α-MnO2 nanowires created abundant channels or pores, the mass transfer of OH was confined and its exposed concentration became higher as the distance towards pore walls became smaller. The heterogeneous spatial distribution of OH avoided the long-ranged diffusion of short-lived free radicals and promoted the reaction kinetic rate within membrane compared to that on particles in bulk solution [14].

    This proposal could be further identified by electron paramagnetic resonance (EPR) and quenching results. From the EPR results (Figs. 5a and b), it was observed that DMPO-OH intensity (Fig. 5b) changed slightly in MnO2-M reactor within 60 min, when most of the confined OH near wall within the membrane were captured. However, the DMPO-OH intensity (Fig. 5a) decreased rapidly in MnO2-P over time due to the fact that most surface generated OH were self-quenched in bulk solution (Eqs. 5 and 6) and were less available for DMPO scavenger.

    (5)

    (6)

    Figure 5

    Figure 5.  (a, b) EPR spectrum at different reaction time of MnO2 particle and MnO2 membrane catalyzed H2O2 system, (c) variances of radical quenching tests under different TBA/H2O2 ratio, (d) relationship between adsorption capacity and reactive rate constant for several organic pollutants.

    TBA was used as a quenching probe to evaluate the reactive efficiency of OH towards BPA in MnO2-M and MnO2-P [34]. Results showed that TBA exhibited higher quenching efficiency towards BPA removal in MnO2-M/H2O2 system than in MnO2-P/H2O2 system, regardless of TBA adding concentration (Fig. 5c). This phenomenon indicated that OH was more thermodynamically favorable to react with TBA in a confined interspace than in a bulk solution. The OH confined in the interspace of α-MnO2 membrane was concentrated so that it possessed more collision probabilities with contaminants, resulting in enhanced efficiency during the reaction. As shown in Fig. 5d, the kobs of different phenolic pollutants in both MnO2-M/H2O2 (Fig. 4d) and MnO2-P/H2O2 (Fig. S5 in Supporting information) systems were positively correlated with their adsorption efficiencies. Notably, slope of kobs versus adsorption efficiency in MnO2-M/H2O2 was higher than that in MnO2-P/H2O2 system, indicating that the contaminant removal in MnO2-M/H2O2 was more dependent on the reactions occurring around the surface possessing interface space. In a continuous flow reaction, the mass transportation between OH and phenolic compounds was confined in channels or pores, thus enhancing the oxidized kinetic rate [35].

    In summary, our facile method for the preparation of ultralong α-MnO2 nanowire membrane exhibited two following advantages. On the one hand, such easy and environmentally friendly method built a mild redox surrounding and steric hindrance to control the growth of α-MnO2 nanowires. On the other hand, the α-MnO2 membrane obtained by assembling these α-MnO2 nanowires could create abundant channels or pores, triggering the nanoconfined effect to enrich the reactive oxygen species (OH) available for targeted contaminants in Fenton-like reaction. This study not only established a novel method to prepare ultralong α-MnO2 nanowires, but also pointed out a potential application form of nanowire-shaped Fenton catalysts in environmental remediation.

    The authors declare no conflict of interest.

    The support from National Natural Science Foundation of China (Nos. 52000050, 52100024 and 42007115), Postdoctoral Science Foundation of China (Nos. 2019M663245 and 2020M670913), Heilongjiang Postdoctoral Fund (No. LBH-Z20063) and State Key Laboratory of Urban Water Resource and Environment (Harbin Institute of Technology) (Nos. 2021TS22 and QAK202111) are greatly appreciated.

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2022.07.007.


    1. [1]

      K. Chen, G.H. Wang, W.B. Li, et al., Chin. Chem. Lett. 25 (2014) 1455–1460. doi: 10.1016/j.cclet.2014.06.014

    2. [2]

      Z. Shangguan, X. Yuan, L. Jiang, et al., Chin. Chem. Lett. 33 (2022) 4719–4731. doi: 10.1016/j.cclet.2022.01.001

    3. [3]

      Y. Liu, F. Liu, N. Ding, et al., Chin. Chem. Lett. 31 (2020) 2539–2548. doi: 10.1016/j.cclet.2020.03.011

    4. [4]

      Y. Liu, F. Li, Q. Xia, et al., Nanoscale 10 (2018) 4771–4778. doi: 10.1039/C7NR09435C

    5. [5]

      Y. Zhang, C. Liu, B. Xu, F. Qi, W. Chu, Appl. Catal. B: Environ. 199 (2016) 447–457. doi: 10.1016/j.apcatb.2016.06.003

    6. [6]

      Z. Wan, J. Wang, J. Hazard. Mater. 324 (2017) 653–664. doi: 10.1016/j.jhazmat.2016.11.039

    7. [7]

      E. Saputra, S. Muhammad, H. Sun, et al., Environ. Sci. Technol. 47 (2013) 5882–5887. doi: 10.1021/es400878c

    8. [8]

      S. Zhu, S.H. Ho, C. Jin, X. Duan, S. Wang, Environ. Sci. Nano 7 (2020) 368–396. doi: 10.1039/C9EN01250H

    9. [9]

      Y. Liu, M. Tourbin, S. Lachaize, P. Guiraud, Powder Technol. 255 (2014) 149–156. doi: 10.1016/j.powtec.2013.08.025

    10. [10]

      J. Zhao, J. Liu, N. Li, et al., Chem. Eng. J. 304 (2016) 737–746. doi: 10.1016/j.cej.2016.07.003

    11. [11]

      L. Zhao, J. Ma, Z.Z. Sun, X.D. Zhai, Appl. Catal. B: Environ. 83 (2008) 256–264. doi: 10.1016/j.apcatb.2008.02.009

    12. [12]

      B. Lan, L. Yu, T. Lin, et al., ACS Appl. Mater. Interfaces 5 (2013) 7458–7464. doi: 10.1021/am401774r

    13. [13]

      J. Qian, X. Gao, B. Pan, Environ. Sci. Technol. 54 (2020) 8509–8526. doi: 10.1021/acs.est.0c01065

    14. [14]

      S. Zhang, M. Sun, T. Hedtke, et al., Environ. Sci. Technol. 54 (2020) 10868–10875. doi: 10.1021/acs.est.0c02192

    15. [15]

      J. Wang, H. Guo, Z. Yang, Y. Mei, C.Y. Tang, J. Membr. Sci. 525 (2017) 298–303. doi: 10.1016/j.memsci.2016.12.003

    16. [16]

      Y. Xie, Y. Liu, Y. Yao, et al., Chin. Chem. Lett. 33 (2022) 1298–1302. doi: 10.1016/j.cclet.2021.07.055

    17. [17]

      H. Jiang, T. Zhao, J. Ma, C. Yan, C. Li, Chem. Commun. 47 (2011) 1264–1266. doi: 10.1039/C0CC04134C

    18. [18]

      J.W. Liu, H.W. Liang, S.H. Yu, Chem. Rev. 112 (2012) 4770–4799. doi: 10.1021/cr200347w

    19. [19]

      F. Cheng, J. Zhao, W. Song, et al., Inorg. Chem. 45 (2006) 2038–2044. doi: 10.1021/ic051715b

    20. [20]

      S. Rong, P. Zhang, F. Liu, Y. Yang, ACS Catal. 8 (2018) 3435–3446. doi: 10.1021/acscatal.8b00456

    21. [21]

      X. Wang, Y. Li, J. Am. Chem. Soc. 124 (2002) 2880–2881. doi: 10.1021/ja0177105

    22. [22]

      X. Wang, S. Ni, G. Zhou, et al., Mater. Lett. 64 (2010) 1496–1498. doi: 10.1016/j.matlet.2010.04.002

    23. [23]

      X. Wang, L. Dou, L. Yang, J. Yu, B. Ding, J. Hazard. Mater. 324 (2017) 203–212. doi: 10.1016/j.jhazmat.2016.10.050

    24. [24]

      S. Annabella, Nat. Mater. 7 (2008) 613. doi: 10.1038/nmat2241

    25. [25]

      A. Vittadini, A. Selloni, F.P. Rotzinger, M. Grätzel, Phys. Rev. Lett. 81 (1998) 2954–2957. doi: 10.1103/PhysRevLett.81.2954

    26. [26]

      J. Bertsch, A.L. Siemund, F. Kremp, V. Muller, Environ. Microbiol. 18 (2016) 2913–2922. doi: 10.1111/1462-2920.13082

    27. [27]

      G. Cheng, S. Xie, B. Lan, et al., J. Mater. Chem. A 4 (2016) 16462–16468. doi: 10.1039/C6TA04530H

    28. [28]

      Q. Zhang, X. Cheng, X. Feng, et al., J. Mater. Chem. 21 (2011) 462–465.

    29. [29]

      J. Yuan, W.N. Li, S. Gomez, S.L. Suib, J. Am. Chem. Soc. 127 (2005) 14184–14185. doi: 10.1021/ja053463j

    30. [30]

      Y. Chen, G. Zhang, H. Liu, J. Qu, Angew. Chem. Int. Ed. 58 (2019) 8134–8138. doi: 10.1002/anie.201903531

    31. [31]

      X. Ai, Y.H. Li, Y.W. Li, T. Gao, K.G. Zhou, Chin. Chem. Lett. 33 (2022) 2832–2844. doi: 10.1016/j.cclet.2021.10.013

    32. [32]

      J.R. Peller, S.P. Mezyk, W.J. Cooper, Res. Chem. Intermed. 35 (2009) 21–34. doi: 10.1007/s11164-008-0012-6

    33. [33]

      S. Zhang, T. Hedtke, Q. Zhu, et al., Environ. Sci. Technol. 55 (2021) 9266–9275. doi: 10.1021/acs.est.1c01391

    34. [34]

      D. Yuan, C. Zhang, S. Tang, et al., Chin. Chem. Lett. 32 (2021) 3387–3392. doi: 10.1016/j.cclet.2021.04.050

    35. [35]

      Y. Oaki, H. Imai, Angew. Chem. 119 (2007) 5039–5043. doi: 10.1002/ange.200700244

  • Figure 1  (a) XRD pattern, (b) EDS spectrum, (c) SEM images and (d) HR-TEM images and SAED pattern of prepared MnO2 nanowires.

    Figure 2  SEM images of MnO2 nanowires prepared under different conditions: (a) CH3CH2OH as reductive reagent at 140 ℃ for 12 h, (b) CH3CHO as reductive reagent at 140 ℃ for 12 h, (c) CH3CH2OH as reductive reagent and H2SO4 as control reagent at 140 ℃ for 12 h, (d) CH3CH2OH as reductive reagent and CH3COOH as control reagent at 140 ℃ for 12 h.

    Figure 3  Schematic illustration for the formation mechanism of α-MnO2 nanowires.

    Figure 4  (a, b) Digital photo and SEM images of obtained MnO2 membrane, (c) BPA removal efficiency and (d) reactive rate constant for organic pollutants of MnO2 membrane catalyzed H2O2 system. Condition: solution volume, 50 mL; MnO2-M/MnO2-P, 1 mg; H2O2, 50 mg/L; BPA, 2 mg/L; pH, 6.75; for continuous flow reaction (MnO2-M), recycle solution speed, 50 mL/min; for batch reaction, stirring speed, 300 r/min.

    Figure 5  (a, b) EPR spectrum at different reaction time of MnO2 particle and MnO2 membrane catalyzed H2O2 system, (c) variances of radical quenching tests under different TBA/H2O2 ratio, (d) relationship between adsorption capacity and reactive rate constant for several organic pollutants.

  • 加载中
计量
  • PDF下载量:  9
  • 文章访问数:  758
  • HTML全文浏览量:  125
文章相关
  • 发布日期:  2023-03-15
  • 收稿日期:  2022-04-22
  • 接受日期:  2022-07-06
  • 修回日期:  2022-06-19
  • 网络出版日期:  2022-07-08
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