A magnetron sputtered Mo3Si thin film: an efficient electrocatalyst for N2 reduction under ambient conditions

Ting Wang ab, Qian Liu b, Tingshuai Li b, Siyu Lu c, Guang Chen d, Xifeng Shi e, Abdullah M. Asiri f, Yonglan Luo *a, Dongwei Ma *g and Xuping Sun *b
aChemical Synthesis and Pollution Control, Key Laboratory of Sichuan Province, School of Chemistry and Chemical Engineering, China West Normal University, Nanchong 637002, Sichuan, China. E-mail: luoylcwnu@hotmail.com
bInstitute of Fundamental and Frontier Sciences, University of Electronic Science and Technology of China, Chengdu 610054, Sichuan, China. E-mail: xpsun@uestc.edu.cn
cGreen Catalysis Center, College of Chemistry, Zhengzhou University, Zhengzhou 450001, Henan, China
dThe Key Laboratory of Life-Organic Analysis, Key Laboratory of Pharmaceutical Intermediates and Analysis of Natural Medicine, School of Chemistry and Chemical Engineering, Qufu Normal University, Qufu 273165, Shandong, China
eCollege of Chemistry, Chemical Engineering and Materials Science, Shandong Normal University, Jinan 250014, Shandong, China
fChemistry Department, Faculty of Science & Center of Excellence for Advanced Materials Research, King Abdulaziz University, P.O. Box 80203, Jeddah 21589, Saudi Arabia
gKey Laboratory for Special Functional Materials of Ministry of Education, School of Materials Science and Engineering, Henan University, Kaifeng 475004, Henan, China. E-mail: dwmachina@126.com

Received 17th November 2020 , Accepted 9th December 2020

First published on 23rd December 2020


Abstract

Industrially, large-scale NH3 production mainly depends on the Haber–Bosch process, which is accompanied by heavy greenhouse gas emission and serious energy consumption. Electrochemical N2 reduction is considered a sustainable strategy to solve this problem. Herein, we report for the first time that a Mo3Si thin film sputtered on graphite paper is a favorable electrocatalyst for NH3 synthesis under ambient conditions. Electrochemical tests suggest a large NH3 yield rate of 2 × 10−10 mol s−1 cm−2 and a high Faraday efficiency of 6.69% at −0.4 V and −0.3 V vs. a reversible hydrogen electrode, respectively, in 0.1 M Na2SO4. It also demonstrates the high electrochemical and structural stability of such a catalyst as well as excellent selectivity for NH3 generation. Density functional theory calculation reveals that the synergy of the metallic conductivity of Mo3Si and the high chemical activity of the exposed Mo ions benefits the adsorption and activation of N2, and a further proton–electron transfer reaction to produce NH3.


As a significant chemical raw material, NH3 is not only widely applied for the fabrication of resin, dyes, and explosives, but also considered as an attractive carbon-neutral energy carrier (17.7 wt% hydrogen content and 3 kW h kg−1 energy density).1–3 Currently, the industrial-scale technology of artificial synthetic NH3 mainly depends on the Haber–Bosch process, which operates under harsh conditions, leading to serious carbon emission and great energy consumption.4 Electrochemical N2 reduction provides an alternative route for NH3 synthesis in an environmentally benign and sustainable manner, but fairly the inertness of the strong N[triple bond, length as m-dash]N triple bonds makes the reaction of N2 fixation still a challenge.5–11 Although a variety of noble metal-based electrocatalysts (Ag,12 Au,13,14 Pd,15,16 and Ru17) perform effectively to drive the N2 reduction reaction (NRR), their large-scale practical applications are still restricted due to their scarcity and expensiveness. As such, it is of great importance to design and develop earth-abundant alternatives.18–36

In nature, molybdenum (Mo) is involved in biological N2 fixation with nitrogenase catalysis under mild conditions.37–39 Additionally, Mo has also been proven to be a promising metal for the homogeneous NH3 functionalization reaction.40–45 However, It is challenging to graft such a catalyst onto electrodes for electrochemical tests.46 Mo-based heterogeneous electrocatalysts can solve this issue, including Mo nanofilms,47 MoS2,48 Mo2C,49 MoO3,50 MoN,51etc. Despite achieving some progress, the search and identification of new Mo-based heterogeneous electrocatalysts for efficient N2-to-NH3 fixation is still worth exploring.

Transition metal silicides as intermetallic compounds have specific crystal and electronic structures, which are different from those of their component metals.52 As new catalytic materials, metal silicides have attracted growing attention by virtue of their unique physical and chemical properties, such as high thermal stability and electrical conductivity,53,54 but their use for electrocatalytic N2 reduction has not been reported before. Herein, we report on the development of a Mo3Si thin film on graphite paper (Mo3Si/GP) by direct current magnetron sputtering for the NRR. When tested in 0.1 M Na2SO4, the resulting Mo3Si/GP achieves a large NH3 yield rate of 2 × 10−10 mol s−1 cm−2 and a high Faraday efficiency (FE) of 6.69% at −0.4 V and −0.3 V vs. a reversible hydrogen electrode (RHE), respectively. Meantime, it also shows high electrochemical and structural stability. The possible NRR catalytic mechanism is also studied by density functional theory (DFT) calculations.

The X-ray diffraction (XRD) pattern of Mo3Si in Fig. 1a shows two obvious diffraction peaks at 41.34° and 65.78° indexed to the (210) and (222) crystal planes of Mo3Si (PDF # 97-064-4417), respectively. The scanning electron microscopy (SEM) image shows the formation of a Mo3Si film on GP (Fig. 1b). Fig. S1 shows the cross-sectional SEM image of Mo3Si, demonstrating that this film is thin in nature with a thickness of about 800 nm. As shown in Fig. 1d, the high-resolution transmission electron microscopy (HRTEM) image taken from the product (Fig. 1c) reveals an interplanar spacing of 0.214 nm indexed to the (210) plane of Mo3Si. The crystalline nature of the Mo3Si thin film is further supported by the selected area electron diffraction (SAED) pattern as shown in Fig. 1e. The scanning TEM (STEM) (Fig. 1f) and corresponding energy-dispersive X-ray (EDX) elemental mapping images conclude the uniform distribution of Mo and Si elements in the Mo3Si thin film. Fig. 1g and h show the X-ray photoelectron spectroscopy (XPS) spectra of Mo3Si in the Mo 3d and Si 2p regions, respectively. For Mo 3d, the banding energy (BE) peaks at 228.2 and 231.5 eV correspond to Mo 3d5/2 and Mo 3d3/2, respectively.55 For Si 2p, the BE peak at 97.7 eV can be attributed to the Mo–Si bond while the other peak at 101.6 eV is associated with the Si–Si bond.56


image file: d0ta11231c-f1.tif
Fig. 1 (a) XRD pattern and (b) SEM image of Mo3Si/GP. (c) TEM image, (d) HRTEM image, and (e) SAED pattern of Mo3Si. (f) STEM and corresponding EDX elemental mapping images of Mo and Si of Mo3Si. XPS spectra of Mo3Si in the (g) Mo 3d and (h) Si 2p regions.

The electrocatalytic NRR performance of Mo3Si/GP (Mo3Si mass loading: 0.4 mg cm−2) was tested in 0.1 M Na2SO4 using a two-compartment electrochemical cell separated by a piece of Nafion 117 membrane, in which Mo3Si/GP, Ag/AgCl, and a graphite rod were used as the working electrode, reference electrode, and counter electrode, respectively. All potentials were calibrated to an RHE scale. The produced NH3 was spectrophotometrically determined by the indophenol blue method,57 and the possible by-product N2H4 was detected by spectrophotometry using the Watt and Chrisp method.58 Their detecting calibration curves are displayed in Fig. S2 and S3. The linear sweep voltammetry curves in a N2-and Ar-saturated 0.1 M Na2SO4 electrolyte (Fig. S5) were used to estimate the range of potentials of Mo3Si/GP for the NRR. Fig. 2a exhibits the time-dependent current density curves of Mo3Si/GP at a series of potentials from −0.3 V to −0.7 V. After electro-measurements, the electrolytes were colored with the indophenol indicator in a light-shielded circumstance for 1 h and the UV-Vis absorption spectra of the electrolytes are shown in Fig. 2b. Its average NH3 yield rate and the corresponding FEs at various potentials are plotted in Fig. 2c, which suggests that Mo3Si/GP attains the largest NH3 yield rate of 2 × 10−10 mol s−1 cm−2 and the highest FE of 6.69% at −0.4 V and −0.3 V, respectively, outperforming most reported NRR electrocatalysts listed in Table S1. We also used ion chromatography to provide a reliable result in Fig. S5 for the quantitative of NH3. When the potential becomes more negative, the HER becomes dominant, resulting in a decreased NH3 yield rate and FE for the NRR.59,60 Meanwhile, the possible by-product N2H4 is not detected (Fig. S6), reflecting the excellent selectivity of Mo3Si/GP for NH3 synthesis. As shown in Fig. 2d, GP obtains a small amount of NH3 (0.6 μg), while Mo3Si/GP acquires a remarkable amount of NH3 (6.1 μg) under the same conditions, suggesting that Mo3Si/GP is highly active for the NRR.


image file: d0ta11231c-f2.tif
Fig. 2 (a) Time-dependent current density curves of Mo3Si/GP at different potentials in 0.1 M Na2SO4. (b) UV-Vis absorption spectra of the electrolytes stained with the indophenol indicator after electrolysis at a series of potentials for 2 h. (c) Corresponding average NH3 yield rate and FEs of Mo3Si/GP at all potentials. (d) Amount of NH3 produced for Mo3Si/GP and GP after charging at −0.4 V for 2 h in 0.1 M Na2SO4.

To investigate the origin of the produced NH3, we also performed a series of control experiments. As shown in Fig. 3a, the experiments were conducted in an Ar-saturated 0.1 M Na2SO4 electrolyte at −0.4 V and a N2-saturated 0.1 M Na2SO4 electrolyte at open-circuit potential. As expected, almost no NH3 is detected and the corresponding UV-Vis absorption spectra are exhibited in Fig. S7. Moreover, we performed the alternative N2-saturated and Ar-saturated tests at −0.4 V for 2 h in 0.1 M Na2SO4. As shown in Fig. 3b, the amount of NH3 detected in the Ar-saturated electrolyte is negligible compared to that in the N2-saturated electrolyte. Both results indicate that the NH3 products are mainly generated from Mo3Si/GP electrocatalysis. Besides, stability is another crucial parameter for electrocatalysts. As shown in Fig. 3c, after 6 consecutive cycle tests, both the NH3 yield rate and FEs exhibit almost no obvious decline. The corresponding time-dependent current density curves and UV-Vis absorption spectra are displayed in Fig. S8. Besides, this catalyst can maintain its catalytic activity for at least 26 h (Fig. 3d). Almost no ammonia was detected in the anodic electrolytes stained with the indophenol indicator (Fig. S9). Furthermore, the photographs of pH test papers of the 0.1 M Na2SO4 aqueous solution before and after 2 h electrolysis indicate that there is almost no obvious change of pH in our experiment (Fig. S10). These data suggest the excellent electrochemical durability of Mo3Si/GP. After a long-term NRR, the XRD analysis confirms that this catalyst is still Mo3Si/GP in nature (Fig. S11), and the SEM image shows that this catalyst also maintains its thin film morphology (Fig. S12). As shown in Fig. S13, after electrolysis, there is no obvious change in the XPS spectra of Mo3Si in the Mo 3d and Si 2p regions.


image file: d0ta11231c-f3.tif
Fig. 3 (a) The amount of NH3 produced for Mo3Si/GP under different electrochemical conditions. (b) NH3 yield rate of alternative N2-saturated and Ar-saturated tests at −0.4 V for 2 h. (c) NH3 yield rate and FE of cycling tests at −0.4 V. (d) Time-dependent current density curve of Mo3Si/GP at −0.4 V for 26 h.

To gain fundamental insights into the NRR mechanism of the Mo3Si catalyst, first-principles calculations based on DFT have been performed, and the relevant computational details are presented in the ESI. The unit cell of Mo3Si and its density of states (DOS) are presented in Fig. S14. The calculated lattice parameter of 4.84 Å (Fig. S14a) is in good agreement with the experimental value of 4.91 Å.61 The plenty electronic states around the Fermi level (Fig. S14b) show the metallic conductivity of Mo3Si, which facilitates the electron transfer during the NRR process. To be specific, the Mo3Si (210) surfaces have been studied for the NRR, which is the most abundant according to the XRD and HRTEM characterization. The Mo3Si (210) (2 × 1) supercell and its DOS are shown in Fig. S15. As shown in Fig. S15a, various potential adsorption sites for N2 molecules have been considered. The obtained stable adsorption configurations are displayed in Fig. S16. It is noted that for all the adsorption configurations the N2 molecules are chemisorbed with significantly negative adsorption energies, ranging from −1.55 to −0.71 eV, and the Mo atoms of the support serve as the active sites. Especially, the configurations in Fig. S16a and b, denoted as Nα2 and Nβ2, respectively, are much more thermodynamically favorable than others, for both of which the N2 molecules are bonded with three Mo atoms. Therefore, in the following we studied the NRR on the Mo3Si (210) surface for the Nα2 and Nβ2 configurations.

The calculated NRR free energy diagrams together with the charge density differences for the Nα2 and Nβ2 configurations are presented in Fig. 4a and b, respectively. The reaction intermediates are displayed in Fig. S17 and S18. The charge density differences in the insets of Fig. S14 suggest that N2 adsorption on the Mo3Si (210) surface follows the “acceptance-donation” mechanism. It is noted that, due to the high chemical activity of Mo, N2 can be strongly activated, as indicated by the significant elongation of the N–N bond (larger than 1.2 Å), and the significant transfer of electrons from the support to the N2 (larger than 1e). There are several well-established reaction pathways for the NRR depending on the specific adsorption modes of N2 molecules.62–64 For the side-on configurations herein, the NRR can proceed through a consecutive or an enzymatic pathway. Thanks to the strong activation of N2, its hydrogenation needs a tiny energy, 0.10 and 0.04 eV, respectively, for Nα2 and Nβ2 configurations. Furthermore, it is found that the enzymatic pathway is energetically much more favorable than the consecutive one, given that for the second hydrogenation step *NH–*NH is more stable than *N–*NH2 by ∼0.5 eV. Therefore, we only studied the enzymatic pathway, as presented in Fig. S14. Importantly, for both Nα2 and Nβ2 configurations, the potential-determining step is the last hydrogenation step to produce the second NH3 molecule, giving a limiting potential of −0.43 V, close to the experimentally optimal working potential (−0.3 or −0.4 V). In addition, H adsorption has been studied to evaluate the NRR selectivity against the HER. From Fig. S19, the active sites for the NRR can also effectively bind H atoms. However, compared with the N2 adsorption, the H adsorption is less stable by more than 0.5 eV. Thus, the Mo3Si (210) surface prefers to adsorb N2 molecules rather than H, which provides a high NRR selectivity, in good agreement with the high FE observed experimentally. Overall, the calculation results indicate that the synergy of the metallic conductivity of Mo3Si and the high chemical activity of Mo ions plays an important role in the adsorption and activation of N2, and a further proton–electron transfer reaction to produce NH3.


image file: d0ta11231c-f4.tif
Fig. 4 The free energy (in eV) diagrams for the NRR on the Mo3Si (210) (2 × 1) surface at zero potential through the enzymatic pathway. (a) and (b) are for Nα2 and Nβ2 configurations presented in Fig. S16a and b, respectively. The calculated charge density difference for Nα2 and Nβ2 are presented as insets, where red and green shadows represent the charge accumulation and loss, respectively. The positions of the N atoms are roughly marked. The blue and mauve spheres denote the Si and Mo atoms, respectively.

In summary, Mo3Si has been proven as a stable and active NRR electrocatalyst with a large NH3 yield rate of 2 × 10−10 mol s−1 cm−2 and a high FE of 6.69% at −0.4 V and −0.3 V vs. RHE, respectively, in neutral media. This work not only provides us an attractive earth-abundant catalyst material for NH3 electrosynthesis under ambient conditions, but would also open up a new avenue to fabricate catalyst films via magnetron sputtering for N2-fixation applications.65

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (No. 22072015).

References

  1. V. Smil, Nature, 1999, 400, 415 CrossRef CAS.
  2. A. Klerke, C. H. Christensen, J. K. Nørskov and T. Vegge, J. Mater. Chem., 2008, 18, 2304–2310 RSC.
  3. R. Schlçgl, Angew. Chem., Int. Ed., 2003, 42, 2004–2008 CrossRef.
  4. I. Dybkjaer, in Ammonia, catalysis and manufacture, ed. A. Nielsen, Springer, Heidelberg, 1995, pp. 199–308 Search PubMed.
  5. C. J. M. van der Ham, M. T. M. Koper and D. G. H. Hetterscheid, Chem. Soc. Rev., 2014, 43, 5183–5191 RSC.
  6. X. Zhu, S. Mou, Q. Peng, Q. Liu, Y. Luo, G. Chen, S. Gao and X. Sun, J. Mater. Chem. A, 2020, 8, 1545–1556 RSC.
  7. B. Ma, H. Zhao, T. Li, Q. Liu, Y. Luo, C. Li, S. Lu, A. M. Asiri, D. Ma and X. Sun, Nano Res., 2020 DOI:10.1007/s12274-020-3049-5.
  8. V. Kyriakou, I. Garagounis, E. Vasileiou, A. Vourros and M. Stoukides, Catal. Today, 2017, 286, 2–13 CrossRef CAS.
  9. C. Guo, J. Ran, A. Vasileff and S. Qiao, Energy Environ. Sci., 2018, 11, 45–56 RSC.
  10. M. D. Fryzuk, J. B. Love, S. J. Rettig and V. G. Young, Science, 1997, 275, 1445–1447 CrossRef CAS.
  11. H. Jia and E. A. Quadrelli, Chem. Soc. Rev., 2014, 43, 547–564 RSC.
  12. H. Huang, L. Xia, X. Shi, A. M. Asiri and X. Sun, Chem. Commun., 2018, 54, 11427–11430 RSC.
  13. S. Li, D. Bao, M. Shi, B. Wulan, J. Yan and Q. Jiang, Adv. Mater., 2017, 29, 1700001 CrossRef.
  14. M. Nazemi, S. R. Panikkanvalappil and M. A. El-Sayed, Nano Energy, 2018, 49, 316–323 CrossRef CAS.
  15. J. Wang, L. Yu, L. Hu, G. Chen, H. Xin and X. Feng, Nat. Commun., 2018, 9, 1795 CrossRef.
  16. G. Deng, T. Wang, A. A. Alshehri, K. A. Alzahrani, Y. Wang, H. Ye, Y. Luo and X. Sun, J. Mater. Chem. A, 2019, 7, 21674–21677 RSC.
  17. Z. Geng, Y. Liu, X. Kong, P. Li, K. Li, Z. Liu, J. Du, M. Shu, R. Si and J. Zeng, Adv. Mater., 2018, 30, 1803498 CrossRef.
  18. T. Wu, H. Zhao, X. Zhu, Z. Xing, Q. Liu, T. Liu, S. Gao, S. Lu, G. Chen, A. M. Asiri, Y. Zhang and X. Sun, Adv. Mater., 2020, 32, 2000299 CrossRef CAS.
  19. T. Xu, D. Ma, T. Li, L. Yue, Y. Luo, S. Lu, A. M. Asiri, C. Yang and X. Sun, Chem. Commun., 2020, 56, 14031–14034 RSC.
  20. S. Gao, Y. Zhu, Y. Chen, M. Tian, Y. Yang, T. Jiang and Z. Wang, Mater. Today, 2019, 28, 17–24 CrossRef CAS.
  21. P. Wei, Q. Geng, A. I. Channa, X. Tong, Y. Luo, S. Lu, G. Chen, S. Gao, Z. Wang and X. Sun, Nano Res., 2020, 13, 2967–2972 CrossRef CAS.
  22. L. Li, C. Tang, B. Xia, H. Jin, Y. Zheng and S. Qiao, ACS Catal., 2019, 9, 2902–2908 CrossRef CAS.
  23. Y. Liu, M. Han, Q. Xiong, S. Zhang, C. Zhao, W. Gong, G. Wang, H. Zhang and H. Zhao, Adv. Energy Mater., 2019, 9, 1803935 CrossRef.
  24. C. Li, J. Yu, L. Yang, J. Zhao, W. Kong, T. Wang, A. M. Asiri, Q. Li and X. Sun, Inorg. Chem., 2019, 58, 9597–9601 CrossRef CAS.
  25. L. Zhang, L. Ding, G. Chen, X. Yang and H. Wang, Angew. Chem., Int. Ed., 2019, 58, 2612–2616 CrossRef CAS.
  26. C. Li, D. Ma, S. Mou, Y. Luo, B. Ma, S. Lu, G. Cui, Q. Li, Q. Liu and X. Sun, J. Energy Chem., 2020, 50, 402–408 CrossRef.
  27. Y. Wang, M. Shi, D. Bao, F. Meng, Q. Zhang, Y. Zhou, K. Liu, Y. Zhang, J. Wang, Z. Chen, D. Liu, Z. Jiang, M. Luo, L. Gu, Q. Zhang, X. Cao, Y. Yao, M. Shao, Y. Zhang, X. Zhang, J. Chen, J. Yan and Q. Jiang, Angew. Chem., Int. Ed., 2019, 58, 9464–9469 CrossRef CAS.
  28. X. Lv, F. Wang, J. Du, Q. Liu, Y. Luo, S. Lu, G. Chen, S. Gao, B. Zheng and X. Sun, Sustainable Energy Fuels, 2020, 4, 4469–4472 RSC.
  29. C. Lv, C. Yan, G. Chen, Y. Ding, J. Sun, Y. Zhou and G. Yu, Angew. Chem., Int. Ed., 2018, 57, 6073–6076 CrossRef CAS.
  30. R. Zhao, Q. Geng, L. Chang, P. Wei, Y. Luo, X. Shi, A. M. Asiri, S. Lu, Z. Wang and X. Sun, Chem. Commun., 2020, 56, 9328–9331 RSC.
  31. Q. Qin, Y. Zhao, M. Schmallegger, T. Heil, J. Schmidt, R. Walczak, G. Demner, H. Jiao and M. Oschatz, Angew. Chem., Int. Ed., 2019, 58, 13101–13106 CrossRef CAS.
  32. T. Wang, S. Li, B. He, X. Zhu, Y. Luo, Q. Liu, T. Li, S. Lu, C. Ye, A. M. Asiri and X. Sun, Chin. J. Catal., 2021, 42, 1024–1029 Search PubMed.
  33. K. Chu, Y. Liu, Y. Li, J. Wang and H. Zhang, ACS Appl. Mater. Interfaces, 2019, 11, 31806–31815 CrossRef CAS.
  34. X. Cheng, J. Wang, W. Xiong, T. Wang, T. Wu, S. Lu, G. Chen, S. Gao, X. Shi, Z. Jiang, X. Niu and X. Sun, ChemNanoMat, 2020, 6, 1315–1319 CrossRef CAS.
  35. L. Huang, J. Wu, P. Han, A. M. Al-Enizi, T. M. Almutairi, L. Zhang and G. Zheng, Small Methods, 2019, 3, 1800386 CrossRef.
  36. J. Gao, X. Lv, F. Wang, Y. Luo, S. Lu, G. Chen, S. Gao, B. Zhong, X. Guo and X. Sun, J. Mater. Chem. A, 2020, 8, 17956–17959 RSC.
  37. B. M. Hoffman, D. Lukoyanov, Z. Yang, D. Dean and L. C. Seefeldt, Chem. Rev., 2014, 114, 4041–4062 CrossRef CAS.
  38. S. F. McWilliams and P. L. Holland, Acc. Chem. Res., 2015, 48, 2059–2065 CrossRef CAS.
  39. L. C. Seefeldt, B. M. Hoffman and D. R. Dean, Annu. Rev. Biochem., 2009, 78, 701–722 CrossRef CAS.
  40. L. A. Wickramasinghe, T. Ogawa, R. R. Schrock and P. Muller, J. Am. Chem. Soc., 2017, 139, 9132–9135 CrossRef CAS.
  41. M. Yu, T. Ogawa, S. R. R. chrock and A. H. Hoveyda, Angew. Chem., Int. Ed., 2015, 54, 215–220 CrossRef CAS.
  42. R. R. Eady, Chem. Rev., 1996, 96, 3013–3030 CrossRef CAS.
  43. R. R. Schrock, Acc. Chem. Res., 2005, 38, 955–962 CrossRef CAS.
  44. H. Tanaka, Y. Nishibayashi and K. Yoshizawa, Acc. Chem. Res., 2016, 49, 987–995 CrossRef CAS.
  45. K. Arashiba, E. Kinoshita, S. Kuriyama, A. Eizawa, K. Nakajima, H. Tanaka, K. Yoshizawa and Y. Nishibayashi, J. Am. Chem. Soc., 2015, 137, 5666–5669 CrossRef CAS.
  46. E. S. Andreiadis, P.-A. Jacques, P. D. Tran, A. Leyris, M. Chavarot-Kerlidou, B. Jousselme, M. Matheron, J. Pécaut, S. Palacin, M. Fontecave and V. Artero, Nat. Chem., 2013, 5, 48–53 CrossRef CAS.
  47. D. Yang, T. Chen and Z. Wang, J. Mater. Chem. A, 2017, 5, 18967–18971 RSC.
  48. Y. Liu, M. Han, Q. Xiong, S. Zhang, C. Zhao, W. Gong, G. Wang, H. Zhang and H. Zhao, Adv. Energy Mater., 2019, 9, 1803935 CrossRef.
  49. H. Cheng, L. Ding, G. Chen, L. Zhang, J. Xue and H. Wang, Adv. Mater., 2018, 30, 1803694 CrossRef.
  50. J. Han, X. Ji, X. Ren, G. Cui, L. Li, F. Xie, H. Wang, B. Li and X. Sun, J. Mater. Chem. A, 2018, 6, 12974–12977 RSC.
  51. L. Zhang, X. Ji, X. Ren, Y. Luo, X. Shi, A. M. Asiri, B. Zheng and X. Sun, ACS Sustainable Chem. Eng., 2018, 6, 9550–9554 CrossRef CAS.
  52. M. E. Schlesinger, Chem. Rev., 1990, 90, 607–628 CrossRef CAS.
  53. S. P. Murarka, Annu. Rev. Mater. Sci., 1983, 13, 117–137 CrossRef CAS.
  54. X. Chen and C. Liang, Catal. Sci. Technol., 2019, 9, 4785–4820 RSC.
  55. X. Zhu, X. Jiang, X. Yao, Y. Leng, L. Wang and Q. Xue, ACS Appl. Mater. Interfaces, 2019, 11, 26880–26890 CrossRef CAS.
  56. C. Deng, Y. Wang, B. He, W. Zhang, F. Dang, H. Wang and Y. Du, J. Mater. Chem. A, 2020, 8, 259–267 RSC.
  57. D. Zhu, L. Zhang, R. E. Ruther and R. J. Hamers, Nat. Mater., 2013, 12, 836–841 CrossRef CAS.
  58. G. W. Watt and J. D. Chrisp, Anal. Chem., 1952, 24, 2006–2008 CrossRef CAS.
  59. X. Ren, J. Zhao, Q. Wei, Y. Ma, H. Guo, Q. Liu, Y. Wang, G. Cui, A. M. Asiri, B. Li, B. Tang and X. Sun, ACS Cent. Sci., 2019, 5, 116–121 CrossRef CAS.
  60. W. Qiu, X. Xie, J. Qiu, W. Fang, R. Liang, X. Ren, X. Ji, G. Cui, A. M. Asiri, G. Cui, B. Tang and X. Sun, Nat. Commun., 2018, 9, 3485 CrossRef.
  61. A. Misra, J. J. Petrovic and T. E. Mitchell, Scr. Mater., 1998, 40, 191–196 CrossRef.
  62. D. Ma, Z. Zeng, L. Liu, X. Huang and Y. Jia, J. Phys. Chem. C, 2019, 123, 19066–19076 CrossRef CAS.
  63. X. Liu, Y. Jiao, Y. Zheng, M. Jaroniec and S.-Z. Qiao, J. Am. Chem. Soc., 2019, 141, 9664–9672 CrossRef CAS.
  64. D. Ma, Z. Zeng, L. Liu and Y. Jia, J. Energy Chem., 2021, 54, 501–509 CrossRef.
  65. F. Wang, H. Zhao, J. Liang, T. Li, Y. Luo, S. Lu, X. Shi, B. Zheng, J. Du and X. Sun, J. Mater. Chem. A, 2020, 8, 20260–20285 RSC.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/d0ta11231c

This journal is © The Royal Society of Chemistry 2021