The effect of hydrogen reduction of α-MnO2 on formaldehyde oxidation: The roles of oxygen vacancies

Liping Zhou Chunying Wang Yaobin Li Xiaofeng Liu Hua Deng Wenpo Shan Hong He

Citation:  Liping Zhou, Chunying Wang, Yaobin Li, Xiaofeng Liu, Hua Deng, Wenpo Shan, Hong He. The effect of hydrogen reduction of α-MnO2 on formaldehyde oxidation: The roles of oxygen vacancies[J]. Chinese Chemical Letters, 2023, 34(3): 107605. doi: 10.1016/j.cclet.2022.06.028 shu

The effect of hydrogen reduction of α-MnO2 on formaldehyde oxidation: The roles of oxygen vacancies

English

  • HCHO is a serious threat to human health in indoor environments, and its main sources are wood furniture and artificial boards [1]. Chronic exposure to HCHO causes a number of diseases such as skin allergies, asthma, acute poisoning, neurasthenia and even a variety of cancers [2]. Therefore, finding an effective way to control indoor formaldehyde is an urgent issue. Several HCHO purification technologies have been developed in recent years, including adsorption [3], photo-catalytic oxidation [4], plasma decomposition with a catalyst [5], and catalytic oxidation [6-8]. Among these technologies, the catalytic oxidation method is an efficient and feasible approach for indoor HCHO elimination without secondary pollution [9].

    Usually, thermal catalytic catalysts include supported noble metal and metal oxide catalysts. Although noble-metal-based catalysts such as Pt- [10-12], Pd- [13-16], Au- [17-19] and Ir-based [20, 21] catalysts exhibit excellent catalytic performance at ambient temperature, their commercial applications are limited by their high cost. On the contrary, metal oxide catalysts such as MnOx-SnO2 [22], MnOx-Co3O4-CeO2 [23], Ag/CeO2 [24], Co3O4 [25], Ce-MnO2 [26], and Cu/MnO2 [27] have attracted more and more attention due to their low price and promising performance. In 2001, Sekine and Nishimura found that MnO2 possessed the best HCHO removal efficiency by comparing various transition metal oxides including CeO2, CuO, TiO2, Fe2O3, Mn3O4 and so on [28]. The past two decades have witnessed the development of manganese dioxide catalysts for the oxidation of HCHO. Manganese dioxide possesses variable valence states and unique physicochemical properties, resulting in superior HCHO catalytic activity compared with other transition metal oxides. Recently, the performance of manganese dioxide in HCHO oxidation has been investigated from many aspects, including crystal structure [29], morphology [30], and crystal facets [31]. Zhang et al. found that δ- and α-MnO2 displayed better HCHO catalyst performance than β- and γ-MnO2 because δ- and α-MnO2 possessed unique tunnel structures and abundant surface lattice oxygen species, which enhanced the HCHO oxidation activity [32]. Chen et al. reported that α-MnO2 possessed a wealth of surface lattice oxygen species, which play a major role in HCHO oxidation [33]. Rong et al. also found that the {310} facets of α-MnO2 promoted the formation of oxygen vacancies, which led to visibly promoted activity for HCHO elimination [34]. α-MnO2, consisting of a tunnel structure with size 0.46 × 0.46 nm, exposes more MnO6 edges, which is conducive to the formation of surface oxygen vacancies [35]. Hence, α-MnO2 is a promising catalyst in the field of indoor HCHO purification. Great efforts have been made to develop methods for improving the catalytic activity of manganese oxides at lower temperatures, including noble metal doping [36, 37], composite material manufacturing [31] and defect adjustment [38]. According to previous works, surface oxygen vacancy formation can enhance the formation of surface-active oxygen species and then improve the performance of MnO2 catalysts for HCHO oxidation [26, 38, 39].

    Therefore, in this work, α-MnO2 catalysts were reduced by hydrogen for different periods of time to control the concentration of oxygen vacancies and then tested for their performance in the catalytic oxidation of HCHO at low temperatures. A dramatic difference in the catalytic activity was clearly observed between the four kinds of α-MnO2. The as-obtained catalyst α-MnO2-2h showed the best catalytic activity among the four catalysts, achieving 100% HCHO conversion at 70 ℃ with a weight hourly space velocity (WHSV) of 60,000 mL g−1 h−1 (Table S1 in Supporting information). The catalysts were carefully characterized by various methods. Based on the results, factors affecting the catalytic activity were elucidated.

    The catalytic performance of α-MnO2-0h, α-MnO2-1h, α-MnO2-2h and α-MnO2-3h in HCHO oxidation was tested, and the results are shown in Fig. 1a. The HCHO conversion of the catalysts at the initial temperature of 30 ℃ were very low, (less than 20%). The HCHO conversion ratio of α-MnO2-0h (30% at 100 ℃) was obviously lower than the other three catalysts, indicating that hydrogen pre-reduction of the α-MnO2 catalyst can effectively enhance its catalytic activity of HCHO oxidation. The temperatures for complete HCHO oxidation over α-MnO2-1h, α-MnO2-2h and α-MnO2-3h catalysts were 90, 70 and 70 ℃, respectively. Meanwhile, as seen in Fig. 1b, there were no other by-products besides CO2. Fig. S1 (Supporting information) shows a comparison of α-MnO2-2h catalyst activities in RH = 0 and RH = 35%. Obviously, the catalytic activity decreased after adding water. T100% of α-MnO2-2h increased from 70 ℃ to 90 ℃. Among these catalysts, α-MnO2-2h showed the best performance of HCHO oxidation, thus the stability of α-MnO2-2h was further evaluated at 70 ℃ (Fig. S2 in Supporting information). The result showed excellent catalytic stability for α-MnO2-2h, maintaining 99% of HCHO conversion for the tested 20 h.

    Figure 1

    Figure 1.  (a) HCHO conversion and (b) CO2 yield of α-MnO2-0h, α-MnO2-1h, α-MnO2-2h and α-MnO2-3h catalysts. Reaction conditions: HCHO concentration = 150 ppm, 20% O2, He balance, and WHSV = 60,000 mL g−1 h−1.

    The N2 adsorption-desorption isotherms of the four catalysts are shown in Fig. S3 (Supporting information). The four catalysts showed typical Ⅳ adsorption isotherms according to the classification criteria proposed by IUPAC, indicating that all of the catalysts were mesoporous materials. Meanwhile, as shown in Table S2 (Supporting information), the specific surface area of the catalyst decreased after hydrogen reduction, indicating that hydrogen reduction may lead to the agglomeration of α-MnO2 particles.

    SEM and TEM images of the MnO2 samples are shown in Fig. S4 (Supporting information). With the increase of hydrogen reduction time, the samples converted from a dispersed sheet structure to solid particles gradually, indicating that the sample particles agglomerated gradually during hydrogen reduction, which was consistent with the results of N2 adsorption-desorption.

    The XRD patterns of the catalysts are shown in Fig. 2. Some peaks were observed for all of the samples at 12.8 (110), 18.1 (200), 28.8 (310), 37.5 (211), 39.0 (330), 60.3 (521), which were assigned to α-MnO2 (JCPDS No. 44-0141) [36, 40]. After prereduction treatment, the XRD patterns of the samples became broader and weaker, indicating the decline of crystallinity. When the sample was reduced for 3 h in hydrogen, diffraction peaks located at 32.4 (103) and 58.5 (321) appeared, which were assigned to Mn3O4 (JCPDS No. 80–0382) [30]. It suggested that MnO2 was partially reduced to Mn3O4 after 3 h hydrogen reduction.

    Figure 2

    Figure 2.  XRD patterns of α-MnO2-0h, α-MnO2-1h, α-MnO2-2h and α-MnO2-3h catalysts.

    The electronic states of manganese on the surface of as-obtained samples were measured by XPS, and the results are show in Fig. 3. The average oxidation state (AOS) of Mn was calculated by following formula [41]: AOS = 8.956 − 1.126 ∆ES, where ∆ES represents the binding energy difference of Mn 3s (Fig. 3a). The AOS of Mn on α-MnO2-0h, α-MnO2-1h, α-MnO2-2h and α-MnO2-3h catalysts were 3.7, 3.5, 3.4 and 3.04, respectively, as shown in Table S2 (Supporting information). The results indicated that the longer the reduction time, the smaller the AOS of Mn, which revealed that MnOx was reduced gradually. According to the results of Mn 2p XPS (Fig. S5 in Supporting information), there were three peaks at 640.4, 641.6 and 643.7 eV for all of the samples, which were attributed to Mn2+, Mn3+ and Mn4+ species [42, 43], respectively. The proportions of Mn3+/MnTotal were calculated, and the results are listed in Table S2 (Supporting information). It is worth noting that the sample of α-MnO2-2h exhibited the largest Mn3+/MnTotal ratio. As we know, oxygen vacancies will be produced when Mn3+ exists in pure manganese dioxide in order to maintain charge balance [44]. That is, α-MnO2-2h possessed the largest content of oxygen vacancies on account of the largest content of Mn3+. The existence of abundant oxygen vacancies on α-MnO2-2h catalyst can enhance the adsorption and activation of molecular oxygen and water and then improve the mobility of surface oxygen species [13, 38, 43, 45], which may be the main reason for the excellent activity of the α-MnO2-2h catalyst for HCHO oxidation. However, when the reduction time further extended to 3 h, Mn2+/MnTotal species increased sharply, which was due to the formation of Mn3O4. Mn3O4 species appeared on α-MnO2-3h catalyst, which may affect the performance of α-MnO2-3h catalyst for HCHO oxidation [46]. As shown in Fig. S6 (Supporting information), compared with Mn2O3 and Mn3O4, the α-MnO2-2h catalyst possessed the best performance of HCHO oxidation.

    Figure 3

    Figure 3.  XPS spectra of α-MnO2-0h, α-MnO2-1h, α-MnO2-2h, and α-MnO2-3h catalysts. (a) Mn 3s. (b) O 1s.

    Fig. 3b shows the XPS spectra of O 1s. The O 1s spectra can be deconvoluted into three peaks to further access the information on intrinsic quality of oxygen species. The peak at the range of 529.3–530 eV can be assigned to lattice oxygen (marked as Olatt); the peak at binding energy 530.7–531.2 eV is assigned to the surface-adsorbed oxygen species (marked as Oads), and the peak at the higher binding energy of 532.2–533.7 eV can be assigned to the surface-adsorbed molecular water (marked as Owat) [25]. It is clearly shown that lattice oxygen and surface-adsorbed oxygen account for a large part of oxygen species on the catalyst surface. Table S2 displays the Oads/Ototal ratio of four as-obtained samples. The Oads/Ototal ratio of the samples followed the order: α-MnO2-2h > α-MnO2-3h > α-MnO2-1h > α-MnO2-0h, in consistent with the order of activity. This result showed that α-MnO2-2h possesses the most abundant surface-adsorbed oxygen. Because α-MnO2-2h possessed sufficient oxygen vacancies, their existence led to oxygen atoms near oxygen vacancies becoming unsaturated, which enhanced the migration of lattice oxygen and acted as surface active oxygen species [47-49].

    Raman and ESR tests were carried out to further illustrate the formation of oxygen vacancies on the catalysts. The Raman spectra of the samples are shown in Fig. 4a. The Raman band at near 650 cm−1 can be assigned to the symmetric stretching vibration Mn—O bonds in the MnO6 octahedron [41, 50, 51]. The Mn-O bond force constant (k) could be calculated by Hooke's law as follows (Eq. 1) [52]:

    (1)

    Figure 4

    Figure 4.  (a) Raman spectra and (b) ESR spectra of α-MnO2-0h, α-MnO2-1h, α-MnO2-2h and α-MnO2-3h catalysts.

    where ω is the Raman shift (cm−1), c is light velocity, and μ is effective mass. According to the formula, there is a positive correlation between ω and the Mn-O bond force constant (k). As shown by the spectra, the band at 656.3 cm−1 for α-MnO2-0h shifts to 650 cm−1 (α-MnO2-1h), 643.8 cm−1 (α-MnO2-2h) and 650 cm−1 (α-MnO2-3h), respectively, after hydrogen reduction for different times. The peak-shift phenomenon indicated the change in the Mn-O bond force among the catalysts. The Mn-O bonds in α-MnO2-2h had the weakest force among the catalysts, which contributed to the production of oxygen vacancies and finally favored HCHO catalytic oxidation [34]. Fig. 4b shows the ESR spectra of the four catalysts. The signal at g ≈ 2.004 was observed for all of the samples, which was ascribed to oxygen vacancies [13, 53]. The signal intensities of oxygen vacancies followed the order of α-MnO2-2h > α-MnO2-3h > α-MnO2-1h > α-MnO2-0h, which was consistent with the results of activity test. It shown that hydrogen reduction weakened the Mn-O bond and increased the content of oxygen vacancies on the catalysts. The presence of abundant oxygen vacancies in α-MnO2-2h facilitated the adsorption, activation, and migration of O2 molecules and further improved the catalytic activity [43, 54].

    The activated oxygen capacities of the samples were further evaluated by H2-TPR, and the results are shown in Fig. S7 (Supporting information). As we know, the peak located in the region below 200 ℃ can be assigned to the surface-activated oxygen species [33]. It is worth noting that the reduction peak of α-MnO2-2h moved to lower temperature in comparison with other samples. In other words, it is more likely to adsorb oxygen molecules from the air to form surface-adsorbed oxygen [55]. The temperature of the reduction peak for α-MnO2-2h was the lowest. It indicated that α-MnO2-2h has superior catalytic capacity, which can facilitate the adsorption and activation of molecular oxygen to form active oxygen species, in consistent with the result of O 1s XPS. Therefore, the α-MnO2-2h catalyst showed the best catalytic performance for HCHO oxidation. The O2-TPD was tested to further enrich the discussion of oxygen activation, which was shown in Fig. S8 (Supporting information). The higher temperature peak at 700 ℃ is attributed to the conversion of bulk lattice oxygen [50]. The desorption peaks of α-MnO2-0h, α-MnO2-1h and α-MnO2-2h remained basically unchanged. However, the desorption peak intensity of α-MnO2-3h is obviously weaker than those of the other samples, indicating the consumption of bulk lattice oxygen in α-MnO2-3h during H2 reduction. This is a good explanation for the presence of the Mn2+ species in α-MnO2-3h.

    In-situ DRIFTS was tested to explore the effect of H2 reduction on reaction intermediates during HCHO oxidation process and to clarify the HCHO oxidation mechanism. As shown in Fig. S9 (Supporting information), 150 ppm of HCHO and O2 were used to pass through the α-MnO2 with the detection of the accumulation of surface reacted intermediates. From the figure we can see that the intermediates, formate species (1375, 1595 cm−1 for ʋ(COO)) [13, 38] appeared in all the samples. However, when the O2 was introduced into the system, the formate species decreased and the hydroxyl species (3677 cm−1 for ʋ(OH)) consumed in α-MnO2-2h evidently. That is, the intermediates conversion rate of α-MnO2-2h was higher than those of the other samples. To sum up, H2 reduction can improve the conversion of intermediates to CO2 and H2O on α-MnO2-2h.

    In summary, the strategy of hydrogen reduction was utilized to obtain α-MnO2 catalysts with different concentrations of oxygen vacancies. The HCHO catalytic oxidation performances of the obtained catalysts followed the order: α-MnO2-2h > α-MnO2-3h > α-MnO2-1h > α-MnO2-0h, and the α-MnO2-2h catalyst could completely oxidize HCHO to CO2 and H2O at 70 ℃. The characterizations indicated that the reduction of the catalyst further generates oxygen vacancies which could enhance the adsorption, activation and mobility of O2 molecules, and thereby enhanced HCHO catalytic oxidation. However, the existence of Mn3O4 decreased the reaction activity. To sum up, the properties and density of oxygen vacancies played a crucial role in the catalytic degradation of HCHO. This work provides a reasonable explanation for this phenomenon and offers a new approach to the design of efficient low-price catalysts for HCHO removal.

    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.

    The work was supported by the Cultivating Project of Strategic Priority Research Program of Chinese Academy of Sciences (No. XDPB1902), the Science and Technology Planning Project of Xiamen City (No. 3502Z20191021), the Science and Technology Innovation "2025″ major program in Ningbo (No. 2022Z028) and Youth Innovation Promotion Association of Chinese Academy of Sciences (No. 2020310).

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2022.06.028.


    1. [1]

      E.L. Hult, H. Willem, P.N. Price, et al., Indoor Air 25 (2015) 523–535. doi: 10.1111/ina.12160

    2. [2]

      T. Salthammer, S. Mentese, R. Marutzky, Chem. Rev. 110 (2010) 2536–2572. doi: 10.1021/cr800399g

    3. [3]

      C. Ma, X. Li, T. Zhu, Carbon 49 (2011) 2869–2877. doi: 10.1016/j.carbon.2011.02.056

    4. [4]

      R. Portela, I. Jansson, S. Suárez, et al., Chem. Eng. J. 310 (2017) 560–570. doi: 10.1016/j.cej.2016.06.018

    5. [5]

      X. Zhu, X. Gao, R. Qin, et al., Appl. Catal. B: Environ. 170 (2015) 293–300. doi: 10.1016/j.apcatb.2015.01.032

    6. [6]

      B. Bai, Q. Qiao, H. Arandiyan, J. Li, J. Hao, Environ. Sci. Technol. 50 (2016) 2635–2640. doi: 10.1021/acs.est.5b03342

    7. [7]

      L. Miao, J. Wang, P. Zhang, Appl. Surf. Sci. 466 (2019) 441–453. doi: 10.1016/j.apsusc.2018.10.031

    8. [8]

      J. Ma, R. Cao, Y. Dang, et al., Chin. Chem. Lett. 32 (2021) 2985–2993. doi: 10.1016/j.cclet.2021.03.031

    9. [9]

      Y. Li, C. Zhang, H. He, J. Zhang, M. Chen, Catal. Sci. Technol. 6 (2016) 2289–2295. doi: 10.1039/C5CY01521A

    10. [10]

      C. Wang, Y. Li, L. Zheng, et al., ACS Catal. 11 (2021) 456–465. doi: 10.1021/acscatal.0c03196

    11. [11]

      W. Ahmad, E. Park, H. Lee, J.Y. Kim, Y. Oh, Appl. Surf. Sci. 538 (2021) 147504. doi: 10.1016/j.apsusc.2020.147504

    12. [12]

      F. Wang, Y. Wang, K. Han, H. Yu, J. Environ. Chem. Eng. 8 (2020) 104041. doi: 10.1016/j.jece.2020.104041

    13. [13]

      C. Wang, Y. Li, C. Zhang, H. He, Appl. Catal. B: Environ. 282 (2021) 119540. doi: 10.1016/j.apcatb.2020.119540

    14. [14]

      K. Li, J. Ji, M. He, H. Huang, Catal. Sci. Technol. 10 (2020) 6257–6265. doi: 10.1039/D0CY01085E

    15. [15]

      Y. Li, C. Wang, C. Zhang, H. He, Top. Catal. 63 (2020) 810–816. doi: 10.1007/s11244-020-01349-1

    16. [16]

      S. Kang, M. Wang, N. Zhu, et al., Chin. Chem. Lett. 30 (2019) 1450–1454. doi: 10.1016/j.cclet.2019.03.023

    17. [17]

      B. Chen, X. Zhu, Y. Wang, L. Yu, C. Shi, Chin. J. Catal. 37 (2016) 1729–1737. doi: 10.1016/S1872-2067(16)62470-1

    18. [18]

      Z. Tang, W. Zhang, Y. Li, et al., Appl. Surf. Sci. 364 (2016) 75–80. doi: 10.1016/j.apsusc.2015.12.112

    19. [19]

      B. Chen, X. Zhu, Y. Wang, et al., Catal. Today 281 (2017) 512–519. doi: 10.1016/j.cattod.2016.06.023

    20. [20]

      Y. Li, X. Chen, C. Wang, C. Zhang, H. He, ACS Catal. 8 (2018) 11377–11385. doi: 10.1021/acscatal.8b03026

    21. [21]

      J. Guo, C. Lin, C. Jiang, P. Zhang, Appl. Surf. Sci. 475 (2019) 237–255. doi: 10.1016/j.apsusc.2018.12.238

    22. [22]

      Y. Wen, T. Xin, J. Li, et al., Catal. Commun. 10 (2009) 1157–1160. doi: 10.1016/j.catcom.2008.12.033

    23. [23]

      S. Lu, K. Li, F. Huang, C. Chen, B. Sun, Appl. Surf. Sci. 400 (2017) 277–282. doi: 10.1016/j.apsusc.2016.12.207

    24. [24]

      L. Ma, D. Wang, J. Li, et al., Appl. Catal. B: Environ. 148-149 (2014) 36–43. doi: 10.1016/j.apcatb.2013.10.039

    25. [25]

      J. Chen, M. Huang, W. Chen, H. Tang, R. Wang, Ind. Eng. Chem. Res. 59 (2020) 18781–18789. doi: 10.1021/acs.iecr.0c03459

    26. [26]

      Z. Lin, S. Rong, P. Zhang, J. Wang, Appl. Catal. B: Environ. 211 (2017) 212–221. doi: 10.1016/j.apcatb.2017.04.025

    27. [27]

      J. Pei, H. Xu, L. Yi, Build. Environ. 84 (2015) 134–141. doi: 10.1016/j.buildenv.2014.11.002

    28. [28]

      Y. Sekine, A. Nishimura, Atmos. Environ. 35 (2001) 2001–2007. doi: 10.1016/S1352-2310(00)00465-9

    29. [29]

      H. Chen, J. He, C. Zhang, H. He, J. Phys. Chem. C 111 (2007) 18033–18038. doi: 10.1021/jp076113n

    30. [30]

      L. Qi, Y. Le, C. Wang, R. Lei, T. Wu, New J. Chem. 45 (2021) 3960–3968. doi: 10.1039/D0NJ05515H

    31. [31]

      H. Li, J. Cao, D. Park, Y. Huang, Environ. Sci. Technol. 53 (2019) 10906–10916. doi: 10.1021/acs.est.9b03197

    32. [32]

      C. Zhang, J. Zhang, L. Wang, Y. Li, H. He, Catal. Sci. Technol. 5 (2015) 2305–2313. doi: 10.1039/C4CY01461H

    33. [33]

      B. Chen, B. Wu, L. Yu, M. Crocker, C. Shi, ACS Catal. 10 (2020) 6176–6187. doi: 10.1021/acscatal.0c00459

    34. [34]

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

    35. [35]

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

    36. [36]

      H. Deng, S. Kang, J. Ma, C. Zhang, H. He, Appl. Catal. B: Environ. 239 (2018) 214–222. doi: 10.1016/j.apcatb.2018.08.006

    37. [37]

      J. Wang, D. Li, P. Zhang, Q. Xu, J. Yu, RSC Adv. 5 (2015) 100434–100442. doi: 10.1039/C5RA17018D

    38. [38]

      J. Wang, J. Li, C. Jiang, et al., Appl. Catal. B: Environ. 204 (2017) 47–155.

    39. [39]

      J. Wang, G. Zhang, P. Zhang, J. Mater. Chem. A 5 (2017) 5719–5725. doi: 10.1039/C6TA09793F

    40. [40]

      C.M. Julien, M. Massot, C. Poinsignon, Spectrochim. Acta A 60 (2004) 689–700. doi: 10.1016/S1386-1425(03)00279-8

    41. [41]

      A. Yusuf, Y. Sun, C. Snape, J. He, C. Wang, Mol. Catal. 497 (2020) 111204. doi: 10.1016/j.mcat.2020.111204

    42. [42]

      B. Bai, Q. Qiao, J. Li, J. Hao, Chin. J. Catal. 37 (2016) 27–31. doi: 10.1016/S1872-2067(15)61026-9

    43. [43]

      Y. Huang, Y. Liu, W. Wang, J. Cao, et al., Appl. Catal. B: Environ. 278 (2020) 119294. doi: 10.1016/j.apcatb.2020.119294

    44. [44]

      X. Liu, K. Zhou, L. Wang, B. Wang, Y. Li, J. Am. Chem. Soc. 131 (2009) 3140–3141. doi: 10.1021/ja808433d

    45. [45]

      H. Zhao, J. Tang, Z. Li, et al., Catal. Sci. Technol. 11 (2021) 7110–7124. doi: 10.1039/D1CY01490K

    46. [46]

      L. Zhang, S. Wang, L. Lv, Y. Ding, S. Wang, Langmuir 37 (2021) 1410–1419. doi: 10.1021/acs.langmuir.0c02841

    47. [47]

      P. Hu, X. Tang, Z. Huang, F. Xu, A. Zakariae, Environ. Sci. Technol. 49 (2015) 2384–2390. doi: 10.1021/es504570n

    48. [48]

      X. Zeng, C. Shan, M. Sun, et al., Chin. Chem. Lett. 33 (2022) 4771–4775. doi: 10.1016/j.cclet.2021.12.085

    49. [49]

      L. Li, X. Feng, N. Yao, S. Chen, M. Xia, ACS Catal. 5 (2015) 4825–4832. doi: 10.1021/acscatal.5b00320

    50. [50]

      X. Li, J. Ma, L. Yang, et al., Environ. Sci. Technol. 52 (2018) 12685–12696. doi: 10.1021/acs.est.8b04294

    51. [51]

      G. Zhu, J. Zhu, W. Jiang, et al., Appl. Catal. B: Environ. 209 (2017) 729–737. doi: 10.1016/j.apcatb.2017.02.068

    52. [52]

      Y. Wei, L. Yan, C. Wang, X. Xu, G. Chen, J. Phys. Chem. B 108 (2004) 18547–18551. doi: 10.1021/jp0479522

    53. [53]

      H. Li, F. Ren, J. Liu, et al., Appl. Catal. B: Environ. 172-173 (2015) 37–45. doi: 10.1007/s40333-014-0072-y

    54. [54]

      L. Li, T. Hua, J. He, Q. Yang, J. Phys. Chem. C 120 (2016) 23660–23668. doi: 10.1021/acs.jpcc.6b08312

    55. [55]

      Y. Wang, K. Liu, J. Wu, Z. Hu, L. Guo, ACS Catal. 10 (2020) 10021–10031. doi: 10.1021/acscatal.0c02310

  • Figure 1  (a) HCHO conversion and (b) CO2 yield of α-MnO2-0h, α-MnO2-1h, α-MnO2-2h and α-MnO2-3h catalysts. Reaction conditions: HCHO concentration = 150 ppm, 20% O2, He balance, and WHSV = 60,000 mL g−1 h−1.

    Figure 2  XRD patterns of α-MnO2-0h, α-MnO2-1h, α-MnO2-2h and α-MnO2-3h catalysts.

    Figure 3  XPS spectra of α-MnO2-0h, α-MnO2-1h, α-MnO2-2h, and α-MnO2-3h catalysts. (a) Mn 3s. (b) O 1s.

    Figure 4  (a) Raman spectra and (b) ESR spectra of α-MnO2-0h, α-MnO2-1h, α-MnO2-2h and α-MnO2-3h catalysts.

  • 加载中
计量
  • PDF下载量:  7
  • 文章访问数:  464
  • HTML全文浏览量:  73
文章相关
  • 发布日期:  2023-03-15
  • 收稿日期:  2022-02-15
  • 接受日期:  2022-06-11
  • 修回日期:  2022-04-13
  • 网络出版日期:  2022-06-16
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