Next Article in Journal
Recent Advances in the Electroreduction of CO2 over Heteroatom-Doped Carbon Materials
Next Article in Special Issue
Accelerating the Oxygen Reduction Reaction and Oxygen Evolution Reaction Activities of N and P Co-Doped Porous Activated Carbon for Li-O2 Batteries
Previous Article in Journal
Gold-Catalyzed Synthetic Strategies towards Four-Carbon Ring Systems
Previous Article in Special Issue
Thermal Profiles of Carbon Fiber Based Anisotropic Thin-Films: An Emerging Heat Management Solution for High-Current Flow Electrocatalysis and Electrochemical Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Simple and Facile Fabrication of Anion-Vacancy-Induced MoO3−X Catalysts for Enhanced Hydrogen Evolution Activity

1
Division of Physics and Semiconductor Science, Dongguk University-Seoul, Seoul 04620, Korea
2
Department of Energy Systems, Soonchunhyang University, Asan 31538, Korea
3
School of Materials Science and Engineering, Kookmin University, Seoul 02707, Korea
*
Authors to whom correspondence should be addressed.
Submission received: 30 September 2020 / Revised: 12 October 2020 / Accepted: 12 October 2020 / Published: 14 October 2020

Abstract

:
Advanced catalysts for clean hydrogen generation and storage offer an attractive possibility for developing a sustainable and ecofriendly future energy system. Transition metal oxides (TMO) are appealing candidates to be largely considered as electrode catalysts. However, for practical applications, there are still challenges—the intrinsic catalytic properties of TMOs should be further improved and TMOs should be synthesized by practical routes for cost-effective and scalable production of catalysts. Therefore, finding promising ways to fabricate highly active TMOs with outstanding electrochemical hydrogen evolution performance is required. Here, we present a direct and facile synthetic approach to successfully provide highly efficient MoO3−X catalysts with electrochemically active oxygen vacancies through a one-step thermal activation process on a Mo metal mesh. Variations in the oxidation states of molybdenum oxides can significantly increase the active sites of the catalysts and improve the electrochemical activity, making these oxide compounds suitable for hydrogen evolution reaction (HER). Compared to the bare Mo mesh and fully oxidized Mo (MoO3) electrodes, the fabricated MoO3−X electrode exhibits better electrochemical performance in terms of overpotentials and Tafel slope, as well as the electrochemical 1000 cycling stability, confirming the improved HER performance of MoO3−X. This provides new insight into the simple procedure suitable for the large-production supply.

Graphical Abstract

1. Introduction

The development of new materials for efficient energy generation is of paramount importance in order to meet ever-increasing energy demands and to provide environmentally friendly energy sources. It is expected that hydrogen gas and corresponding energy applications from water-splitting reactions should be capable of providing both high energy capacity and renewable (ecofriendly) resources [1,2,3]. Further research on catalytic electrode materials for an efficient hydrogen evolution reaction (HER) are needed to maximize its catalytic performance. Currently, earth abundant transition metal-based oxides (TMOs) have emerged as promising alternative electrocatalysts for noble metal replacement based on theories and experiments on their metal active sites, electrical structures, and surface properties. Those electrode materials have demonstrated their catalytic efficiency and high catalyst stability during HER regardless of electrolyte pH [4,5,6]. There are still a lot of challenges involved in using these materials as industrial electrode candidates due to their relatively low catalyst performance compared to noble metals, e.g., platinum, and their synthesis methods, which still lack scalability. Enhancing the manufacturing scalability of TMOs and finding their simple fabrication process into high intrinsic catalytic performance are issues that need to be addressed.
To this end, numerous studies are being solicited by introducing and engineering anion defects into the crystal structure of TMOs [7,8,9]. It has been reported that anion deficiency in TMOs can provide favorable electrochemical kinetics as a result of the reduction of transition metal orbitals, which are regarded as the electrochemical active sites to promote gas evolution reactions [10,11,12]. However, it has been noted that, in hydrogen electrode applications, TMOs with anion vacancies cannot maintain their original structure during the adsorption of electrolyte ions while generating hydrogen gas. In other words, failure to accomplish the structural stability of TMO materials will cause a drastic decrease in electrochemical cyclability. Nevertheless, it has been demonstrated that certain structured and layered materials with vacancies can successfully exploit the favorable electrochemical kinetics during the adsorption and desorption of ions [13,14,15]. For example, molybdenum trioxide (MoO3) is one of the promising TMO candidates with layered structures. Among many oxide materials, Molybdenum oxide (MoO3) is an environmentally benign and low-cost chemical compound, which shows HER performance. In particular, its orthorhombic phase with the layered morphology are electrochemically stable due to its unique structure [16,17]. Its catalytic activity and electrochemical properties can be systematically managed through the formation of anion vacancy with rigid structural stability. Specifically, anion deficient MoO3 (MoO3−X) can maintain its original layered characteristics consisting of distorted MoO6 octahedron sites that allow electrolyte ions to favorably intercalate and react with Mo sites [18,19]. Therefore, MoO3−X can be successfully used as the electrode material in lithium-ion batteries and supercapacitors, as well as HER systems. However, synthesizing MoO3−X electrodes in a productive method remains a challenge due to its multistep synthetic process [20,21]. Especially, to induce oxygen vacancies, further hydrogen treatment under high vacuum is required, which cannot be the mass-productive method by using highly explosive H2 gas. Moreover, instead of using Mo precursors, direct synthesis on molybdenum substrate can be crucial to fabricate the electrode in a productive and scalable level. This is because the Mo sources originating from Mo substrates (meshes) and oxygen atoms can be reacted and converted into MoO3−X electrode when the oxygen sources are limited compared to the Mo sources.
Herein, we synthesize the molybdenum oxide electrode materials with a controlled level of oxygen vacancies through a simple and facile thermal activation process for HER electrocatalysts, exhibiting the enhanced electrochemical performance in an alkaline media. Synthesizing catalytic materials directly on its metal substrate, such as the Mo mesh, can be beneficial for efficiently utilizing the corresponding transition metal compounds instead of using active material ink combined with carbon compounds. It has been previously reported that MoO3 can be obtained through the oxidation of a transition metal during annealing at a high temperature in an air atmosphere without additional precursors. Following the annealing process, placing the as-synthesized MoO3 under a reducing atmosphere condition, such as H2 gas, induced MoO3−X [22,23,24]. Here, we successfully and simply produce MoO3−X electrodes with the one-step oxidation process under argon on the Mo mesh. The remarkable material and morphological properties of MoO3−X ensure that the catalytic electrode exhibits a relatively low overpotential with a Tafel slope of 69 mV dec−1 compared to that of MoO3 with a Tafel slope of 91 mV dec−1. Along with its enhanced HER performance, the MoO3−X electrodes show negligible performance degradation even with 1000 cycles of the linear sweep voltammetry (LSV) measurements. This confirms the electrochemical properties of MoO3−X synthesized by our simple and facile one-step oxidation process are comparable to the reported MoO3 electrodes.

2. Results

Promising electrocatalyst candidates with thermodynamically certain energy in water electrolysis should be recognized as ideal candidates with the favorable absorption of hydrogen ions and the desorption of H2 gas in terms of free energy of hydrogen absorption ( G H ° 0) [25]. In this regard, among noble metals, Pt is used as a conventional reference catalyst for HER due to its calculated G H ° close to zero, and it shows high electrical conductivity and outstanding electrochemical performance. Pt/C is an electrocatalyst composed of platinum supported on carbon materials, which can synergistically employ its large surface area and a large number of corresponding active sites. Therefore, the electrochemical activity, electrode conductivity, and large surface area of target materials are critical as an important indicator to access G H ° and electrochemical performance in HER [6].
The MoO3 electrodes with anion vacancy and nanostructured morphology were simply synthesized by the thermal activation process. Firstly, oxygen vacancies in MoO3 can improve the electrical conductivity of the electrode, which is reported in many other articles [26]. Second, the nanostructured morphology of the electrode can provide the large surface area. Figure 1 represents MoO3−X synthesized by controlling the temperature of the activation process without the addition of a conventional reducing gas source, such as hydrogen. A cleaned Mo mesh was introduced at the middle of the thermal annealing quartz tube for thermal treatment under an ambient flow rate of Ar at different temperatures for MoO3-x and MoO3. The MoO3−X electrode was fabricated at low temperature condition (~450 °C), and the MoO3 electrode was synthesized at high temperature condition (~600 °C). Moreover, the enhanced HER performance is reported due to the increased surface area of MoO3−X since its nanostructure can enhance the activity of TMOs in catalytic reactivity. Combining all catalytic features, the as-prepared MoO3−X electrodes can utilize the aforementioned advantages to improve the catalytic activity.
To assess the morphological properties of MoO3 and MoO3−X, various scanning electron microscopy (SEM) analysis techniques were carried out. Figure 2b,c,e,f shows the SEM images of MoO3−X and MoO3, respectively. As can be seen from low-magnification SEM images, the molybdenum trioxide materials were grown and uniformly distributed on the Mo mesh for both the samples. With the higher magnification, the MoO3−X electrode shows the small nanosized grains with a diameter size of 200–400 nm. In contrast, the larger microsized grains were found in the MoO3 electrode. Based on the SEM results, it can be confirmed that MoO3−X, with a smaller particle size than MoO3, will provide the large specific surface area and a large number of the corresponding electrochemically active sites. Finally, Figure 2a,d represents the schematic illustration of both MoO3−X and MoO3 electrodes, respectively. We have used the lack of oxygen conditions to induce the oxygen vacancy in the MoO3−X electrodes. After the low-temperature heat treatment, even though the temperature is high enough to overcome the activation energy of oxidization, there are not enough oxygen sources to react with the Mo substrate. Therefore, oxygen vacancies were generated. During the oxidation process, the adsorption of oxygen on the surface of metal is usually not a rate-limiting step. However, due to the lack of oxygen and the abundance of metal sources, it becomes a rate-limiting step, and the oxygen vacancies of MoO3 occur [27].
In Figure 3a, the color change of MoO3 proves the presence of oxygen vacancies. The MoO3−X electrode has a blue color on the surface of the Mo mesh, while the as-synthesized MoO3 electrode possesses a bright gray and black color after the excess thermal activation process. Similar to the cases with other oxygen-deficient materials such as WO3−x, it has been noted that the blue color of the TMO is induced from characteristic outer d-shell electron density (the delocalization of the electron density) [28,29,30]. In the MoO3−X articles, the presence of oxygen vacancy has been signified through the color change [22]. The blue color of the as-prepared MoO3−X electrode might be the indirect evidence to prove the presence of oxygen vacancy. To further understand the crystal phases and oxidation states, the MoO3 and MoO3−X electrodes were examined by Raman and X-ray photoelectron spectroscopy (XPS) as shown in Figure 3b–d. From the Raman spectroscopy in Figure 3b, three sharp peaks at 280 cm−1, 663 cm−1, and 818 cm−1 are identified and well assigned to the orthorhombic MoO3. Moreover, two relatively distinct peaks at 490 cm−1 and 890 cm−1 correspond to the stretching mode of MoO3−X [10,31]. The surface electronic and oxidation states of the as-prepared samples were characterized by XPS. Figure 3c,d compares the high resolution XPS spectra of the Mo 3d of MoO3−X and MoO3, respectively. The Mo6+ 3d doublet is clearly observed at about 235.9 eV (Mo6+ 3d3/2) and at 232.7 eV (Mo6+ 3d5/2) with a d spin-orbit splitting of 3.13 eV. Moreover, there are two additional doublets at 234.7 eV and 231.5 eV, indicating the oxidation states of Mo5+, which cannot be found in the XPS spectra of MoO3, as shown in Figure 3d [14,15,16]. The various material characterization results based on the SEM, XPS, and Raman spectroscopy demonstrate that the existence of anion deficiency and nanostructured surface morphologies on the electrodes will be beneficial for the overall electrochemical activity. The oxygen vacancy and reduced oxidation states may improve the electrical conductivity of the electrode because of the increased carrier concentration. Moreover, the incorporation of oxygen vacancies can provide more surface active sites and enhance mass transport kinetics [14,21].
The electrocatalytic activity of the as-synthesized MoO3−X and MoO3 electrodes was examined in an alkaline aqueous solution (0.1 M KOH). In all experiments, the three-electrode system with Pt and Ag/AgCl was used as the counter electrode and reference electrode, respectively, while the Mo mesh with MoO3 and MoO3−X was used as the working electrode. In addition, the 20 wt% Pt/C was used as the reference electrocatalyst to objectively figure out the performance of the as-synthesized electrodes. Linear sweep voltammetry (LSV) was performed to measure the electrochemical performance of the electrocatalysts (Figure 4b–d), and to compare the effects of the anion deficiency of the as-synthesized MoO3−X electrodes. Schematically, HER begins through the transfer of electrolyte ions, which are being drawn by the applied potentials on working electrodes (Figure 4a). The polarization curves in Figure 4b show the generated electrons as the estimated current density (mA cm−2), that is obtained by the adsorbed protons. The electrode reaches at a current density of 10 mA cm−2 at a potential of 187 mV for the MoO3-x electrode, while the overpotential of MoO3 at a current density of 10 mA cm−2 is at 261 mV. This overpotential value of MoO3−X is closer to that of the Pt/C catalyst system (56 mV at a current density of 10 mA cm−2) compared to MoO3, indicating the enhanced catalytic activity of MoO3−X. The LSV curves also demonstrate that the significant effects of the existence of anion deficiency also result in the overpotential reduction at a current density of 50 mA cm−2. The comparative Tafel plots and calculated Tafel slopes are derived from the polarization curves. The oxidation state of Mo is strongly related to how well Mo can carry out a series of the corresponding hydrogen reduction and adsorption processes (HER). The performance of the MoO3−X electrodes is improved, but still follows the Heyrovsky mechanism [32,33]. In Figure 4c, the commercial Pt/C sample shows a Tafel slope of 57 mV dec−1. The MoO3 and MoO3−X electrodes show Tafel slopes of 91 mV dec−1 and 69 mV dec−1, respectively, indicating enhancement of HER activity using the MoO3−X electrode. Specifically, the effects of anion deficiency have shown the decrease of the Tafel slope by 22 mV dec−1. In addition to the activity and Tafel slope of the samples, cycling stability is another critical concern for TMO electrocatalysts with anion vacancies. The cycled LSV was conducted for the stability test of MoO3−X and an HER performance test under the same conditions. The cycled LSV was performed from +0.2 V versus reversible hydrogen electrode (RHE) to −0.4 V versus RHE for 1000 cycles with a slight initial increase, but negligible performance deterioration (15 mV and 12 mV increase at current density of 10 mA cm−2 and 50 mA cm−2, respectively) as shown in Figure 4d, confirming the high electrochemical and structural stability of MoO3−X in HER.

3. Materials and Methods

3.1. Preparations of Materials

The structured MoO3−X electrodes were synthesized by the simple thermal oxidation process on the Mo mesh. First, the Mo mesh (Fisher Scientific, Hampton, NH, USA) was ultrasonically cleaned in ethanol, iso-propanol, and deionized water, and blown dry with nitrogen. The Mo mesh was introduced at the middle of the quartz tube, which was placed in the tube furnace. The temperature of the furnace was raised to the reaction temperature of 450–600 °C at a ramp rate of 5 °C/min under an ambient flow rate of Ar at 30 standard cubic centimeters per minute (sccm), and the sample was thermally oxidized for 2 h in an argon atmosphere.

3.2. Material Characterization

The morphologies of the as-prepared electrodes were investigated by scanning electron microscopy (SEM) using Hitachi S-4300 (Hitachi, Oxford, UK). Raman spectroscopy was conducted using Jobin Yvon LabRam Aramis Raman spectroscopy (Oxford, UK) with a 532 nm laser with a power of ∼20 μW and a spot size of 1.09 μm2 under ambient pressure and room temperature. X-ray photoelectron spectroscopy (XPS, Oxford, UK) was performed on a Thermo scientific K-Alpha instrument with an Al Kα X-ray source. The spectrometer was adjusted to align a binding energy of 284.5 eV for the C 1s line.

3.3. Electrochemical Measurements

HER catalytic activity tests were carried out in a standard three-electrode setup controlled by a Metrohm Autolab (Metrohm, Oxford, UK, 2019). A working electrode (Mo oxides with an Mo mesh), an Ag/AgCl (in 1 M KCl, aq) reference electrode, a graphite rod counter electrode, and a 0.1 M KOH solution as the electrolyte were used. The reference electrode was calibrated against and converted to a reversible hydrogen electrode (RHE).

4. Conclusions

The direct and facile synthesis method for the fabrication of the MoO3−X electrode has been validated in this study. The MoO3−X electrode was obtained directly through a thermal activation process on the Mo mesh, which is used as the working electrode with high electrical conductivity and large surface area, as well as the Mo precursor-free synthesis. By applying the activation procedures without the additional Mo metal precursors, the MoO3 and nonstoichiometric MoO3−X electrodes with oxygen vacancies were fabricated on the surface of the Mo mesh. The MoO3−X electrodes showed superior performance in HER. The MoO3−X electrodes require a potential of 187 mV versus RHE to obtain a current density of 10 mA cm−2, which is 74 mV less than the MoO3 electrode and also represents the negligible electrocatalytic degradation during 1000 cycles of LSV. The verification of the effective one-step synthetic method by simply controlling the oxygen levels in TMOs could contribute to their practical applications in the hydrogen catalytic field.

Author Contributions

Conceptualization, S.J. and J.H.; methodology, S.J. and J.H.; validation, S.J., Y.-W.L., and J.H.; formal analysis, S.J. and J.H.; investigation, S.J. and J.H.; resources, S.J. and J.H.; data curation, S.J. and J.H.; writing—original draft preparation, S.J., J.H., and J.I.S.; writing—review and editing, S.J., Y.-W.L., J.H., and J.I.S.; visualization, Y.-W.L., J.H., and J.I.S.; supervision, J.H. and J.I.S.; project administration, J.H. and J.I.S.; funding acquisition, J.H. and J.I.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Research Foundation of Korea (NRF) grant funded by the Korean government (MSIT) (2019R1A2C1007883, 2019M1A2A2065616 and 2020R1F1A1068979).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dresselhaus, M.S.; Thomas, I.L. Alternative energy technologies. Nature 2014, 414, 332–337. [Google Scholar] [CrossRef]
  2. Schlapbach, L. Hydrogen-fuelled vehicles. Nature 2009, 406, 809–811. [Google Scholar] [CrossRef] [PubMed]
  3. Turner, J.A. A realizable renewable energy future. Science 1999, 285, 687–689. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Lu, Q.; Yu, Y.; Ma, Q.; Chen, B.; Zhang, H. 2D transition-metal-dichalcogenide-nanosheet-based composites for photocatalytic and electrocatalytic hydrogen evolution reactions. Adv. Mater. 2016, 28, 1917–1933. [Google Scholar] [CrossRef] [PubMed]
  5. Zhu, Y.P.; Guo, C.; Zheng, Y.; Qiao, S.Z. Surface and interface engineering of noble-metal-free electrocatalysts for efficient energy conversion processes. Acc. Chem. Res. 2017, 50, 915–923. [Google Scholar] [CrossRef]
  6. Zheng, Y.; Jiao, Y.; Jaroniec, M.; Qiao, S.Z. advancing the electrochemistry of the hydrogen evolution reaction through combining experiment and theory. Angew. Chem. 2015, 54, 52–65. [Google Scholar] [CrossRef]
  7. Xu, L.; Jiang, Q.; Xiao, Z.; Li, X.; Huo, J.; Wang, S.; Dai, L. Plasma-engraved Co3O4 nanosheets with oxygen vacancies and high surface area for the oxygen evolution reaction. Angew. Chem. 2016, 55, 5277–5281. [Google Scholar] [CrossRef]
  8. Liu, R.; Wang, Y.; Liu, D.; Zou, Y.; Wang, S. Water-plasma-enabled exfoliation of ultrathin layered double hydroxide nanosheets with multivacancies for water oxidation. Adv. Mater. 2017, 29, 1701546. [Google Scholar] [CrossRef]
  9. Kim, J.; Yin, X.; Tsao, K.-C.; Fang, S.; Yang, H. Ca2Mn2O5 as oxygen-deficient perovskite electrocatalyst for oxygen evolution reaction. J. Am. Chem. Soc. 2014, 136, 14646–14649. [Google Scholar] [CrossRef]
  10. Wu, M.; Ke, S.; Chen, W.; Zhang, S.; Zhu, M.; Zhang, Y.; Foo, M.L.; Tang, L. Optimization of the facet structure of cobalt oxide catalysts for enhanced hydrogen evolution reaction. Catal. Sci. Technol. 2020, 10, 1040–1047. [Google Scholar] [CrossRef]
  11. Lovell, E.; Lu, X.; Zhang, Q.; Scott, J.; Amal, R. From passivation to activation–tunable nickel/nickel oxide for hydrogen evolution electrocatalysis. Chem. Commun. 2020, 56, 1709–1712. [Google Scholar] [CrossRef] [PubMed]
  12. Nayak, A.K.; Verma, M.; Sohn, Y.; Deshpande, P.A.; Pradhan, D. Highly active tungsten oxide nanoplate electrocatalysts for the hydrogen evolution reaction in acidic and near neutral electrolytes. ACS Omega 2017, 2, 7039–7047. [Google Scholar] [CrossRef] [PubMed]
  13. Liu, S.; Chen, C.; Zhang, Y.; Zheng, Q.; Zhang, S.; Mu, X.; Chen, C.; Ma, J.; Mu, S. Vacancy-coordinated hydrogen evolution reaction on MoO3−x anchored atomically dispersed MoRu pairs. J. Mater. Chem. A 2019, 7, 14466–14472. [Google Scholar] [CrossRef]
  14. Kim, H.; Cook, J.B.; Lin, H.; Ko, J.S.; Tolbert, S.H.; Ozolins, V.; Dunn, B. Oxygen vacancies enhance pseudocapacitive charge storage properties of MoO3−x. Nat. Mater. 2017, 16, 454–460. [Google Scholar] [CrossRef] [PubMed]
  15. Shen, S.; Zhang, X.; Cheng, X.; Xu, Y.-M.; Gao, S.; Zhao, H.; Zhou, X.; Huo, L.-H. Oxygen-vacancy-enriched porous α-MoO3 nanosheets for trimethylamine sensing. ACS Appl. Nano Mater. 2019, 2, 8016–8026. [Google Scholar] [CrossRef]
  16. Yu, M.; Shao, H.; Wang, G.; Yang, F.; Liang, C.; Rozier, P.; Wang, C.-Z.; Lu, X.; Simon, P.; Feng, X. Interlayer gap widened α-phase molybdenum trioxide as high-rate anodes for dual-ion-intercalation energy storage devices. Nat. Commun. 2020, 11, 1348. [Google Scholar] [CrossRef] [Green Version]
  17. Brezesinski, T.; Wang, J.; Tolbert, S.H.; Dunn, B. Ordered mesoporous α-MoO3 with iso-oriented nanocrystalline walls for thin-film pseudocapacitors. Nat. Mater. 2010, 9, 146–151. [Google Scholar] [CrossRef]
  18. Yazdani, S.; Kashfi-Sadabad, R.; Huan, T.D.; Morales-Acosta, M.D.; Pettes, M.T. Polyelectrolyte-assisted oxygen vacancies: A new route to defect engineering in molybdenum oxide. Langmuir 2018, 34, 6296–6306. [Google Scholar] [CrossRef]
  19. Shi, X.-R.; Wang, J.; Hermann, K. Theoretical cluster studies on the catalytic sulfidation of MoO3. J. Phys. Chem. C 2010, 114, 6791–6801. [Google Scholar] [CrossRef]
  20. Borgschulte, A.; Sambalova, O.; Delmelle, R.; Jenatsch, S.; Hany, R.; Nüesch, F. Hydrogen reduction of molybdenum oxide at room temperature. Sci. Rep. 2017, 7, 40761. [Google Scholar] [CrossRef] [Green Version]
  21. Wu, Q.-L.; Zhao, S.-X.; Yu, L.; Zheng, X.-X.; Wang, Y.-F.; Yu, L.-Q.; Nan, C.-W.; Cao, G. Oxygen vacancy-enriched MoO3−x nanobelts for asymmetric supercapacitors with excellent room/low temperature performance. J. Mater. Chem. A 2019, 7, 13205–13214. [Google Scholar] [CrossRef]
  22. Luo, Z.; Miao, R.; Huan, T.D.; Mosa, I.M.; Poyraz, A.S.; Zhong, W.; Cloud, J.E.; Kriz, D.A.; Thanneeru, S.; He, J.; et al. Mesoporous MoO3-x material as an efficient electrocatalyst for hydrogen evolution reactions. Adv. Energy Mater. 2016, 6, 1600528. [Google Scholar] [CrossRef]
  23. Liu, J.; Zhu, D.; Zheng, Y.; Vasileff, A.; Qiao, S.Z. Self-supported earth-abundant nanoarrays as efficient and robust electrocatalysts for energy-related reactions. ACS Catal. 2018, 8, 6707–6732. [Google Scholar] [CrossRef]
  24. Zhao, H.; Zhu, Y.-P.; Yuan, Z.-Y. Three-dimensional electrocatalysts for sustainable water splitting reactions. Eur. J. Inorg. Chem. 2016, 2016, 1916–1923. [Google Scholar] [CrossRef]
  25. Zou, X.; Zhang, Y. Noble metal-free hydrogen evolution catalysts for water splitting. Chem. Soc. Rev. 2015, 44, 5148. [Google Scholar] [CrossRef]
  26. Kashfi-Sadabad, R.; Yazdani, S.; Huan, T.D.; Cai, Z.; Pettes, M.T. Role of oxygen vacancy defects in the electrocatalytic activity of substoichiometric molybdenum oxide. J. Phys. Chem. C 2018, 122, 18212–18222. [Google Scholar] [CrossRef]
  27. Oleksak, R.P.; Kapoor, M.; Perea, D.; Holcomb, G.R.; Doğan, Ö.N. The role of metal vacancies during high-temperature oxidation of alloys. NPJ Mater. Degrad. 2018, 2, 25. [Google Scholar] [CrossRef]
  28. Yamashita, T.; Yokoyama, H. Molybdenum anode: A novel electrode for enhanced power generation in microbial fuel cells, identified via extensive screening of metal electrodes. Biotechnol. Biofuels 2018, 11, 39. [Google Scholar] [CrossRef] [Green Version]
  29. Xu, Y.; Zheng, C.; Wang, S.; Hou, Y. 3D arrays of molybdenum sulphide nanosheets on Mo meshes: Efficient electrocatalysts for hydrogen evolution reaction. Electrochim. Acta 2015, 174, 653–659. [Google Scholar] [CrossRef]
  30. Manthiram, K.; Alivisatos, A.P. Tunable localized surface plasmon resonances in tungsten oxide nanocrystals. J. Am. Chem. Soc. 2012, 134, 3995–3998. [Google Scholar] [CrossRef]
  31. Dieterle, M.; Weinberg, G.; Mestl, G. Raman spectroscopy of molybdenum oxides. Phys. Chem. Chem. Phys. 2002, 4, 812–821. [Google Scholar] [CrossRef]
  32. Li, Y.; Wang, H.; Xie, L.; Liang, Y.; Hong, G.; Dai, H. MoS2Nanoparticles grown on graphene: An advanced catalyst for the hydrogen evolution reaction. J. Am. Chem. Soc. 2011, 133, 7296–7299. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Datta, R.S.; Haque, F.; Mohiuddin, M.; Carey, B.J.; Syed, N.; Zavabeti, A.; Zhang, B.; Khan, H.; Berean, K.J.; Ou, J.Z.; et al. Highly active two dimensional α-MoO3−x for the electrocatalytic hydrogen evolution reaction. J. Mater. Chem. A 2017, 5, 24223–24231. [Google Scholar] [CrossRef]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Figure 1. Schematic illustration of the MoO3−X electrodes and their synthesis process.
Figure 1. Schematic illustration of the MoO3−X electrodes and their synthesis process.
Catalysts 10 01180 g001
Figure 2. Schematic illustration of the as-prepared (a) MoO3−X and (d) MoO3 electrodes. Low-magnification scanning electron microscopy (SEM) images of (b) MoO3−X and (e) MoO3. High-magnification SEM images of (c) MoO3−X and (f) MoO3.
Figure 2. Schematic illustration of the as-prepared (a) MoO3−X and (d) MoO3 electrodes. Low-magnification scanning electron microscopy (SEM) images of (b) MoO3−X and (e) MoO3. High-magnification SEM images of (c) MoO3−X and (f) MoO3.
Catalysts 10 01180 g002
Figure 3. (a) Real photo images of the clean Mo mesh, MoO3−X, and MoO3 electrodes. (b) Raman spectra of MoO3−X. Mo 3d X-ray photoelectron spectroscopy (XPS) spectra of (c) MoO3−X and (d) MoO3.
Figure 3. (a) Real photo images of the clean Mo mesh, MoO3−X, and MoO3 electrodes. (b) Raman spectra of MoO3−X. Mo 3d X-ray photoelectron spectroscopy (XPS) spectra of (c) MoO3−X and (d) MoO3.
Catalysts 10 01180 g003
Figure 4. (a) Schematic illustration of the MoO3−X electrode during hydrogen evolution reactions. (b) Polarization curves of the MoO3, MoO3−X, and Pt/C electrodes. (c) Tafel plots of the MoO3, MoO3−X, and Pt/C electrodes. (d) Stability test of the MoO3−X electrode.
Figure 4. (a) Schematic illustration of the MoO3−X electrode during hydrogen evolution reactions. (b) Polarization curves of the MoO3, MoO3−X, and Pt/C electrodes. (c) Tafel plots of the MoO3, MoO3−X, and Pt/C electrodes. (d) Stability test of the MoO3−X electrode.
Catalysts 10 01180 g004

Share and Cite

MDPI and ACS Style

Jo, S.; Lee, Y.-W.; Hong, J.; Sohn, J.I. Simple and Facile Fabrication of Anion-Vacancy-Induced MoO3−X Catalysts for Enhanced Hydrogen Evolution Activity. Catalysts 2020, 10, 1180. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10101180

AMA Style

Jo S, Lee Y-W, Hong J, Sohn JI. Simple and Facile Fabrication of Anion-Vacancy-Induced MoO3−X Catalysts for Enhanced Hydrogen Evolution Activity. Catalysts. 2020; 10(10):1180. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10101180

Chicago/Turabian Style

Jo, Seunghwan, Young-Woo Lee, John Hong, and Jung Inn Sohn. 2020. "Simple and Facile Fabrication of Anion-Vacancy-Induced MoO3−X Catalysts for Enhanced Hydrogen Evolution Activity" Catalysts 10, no. 10: 1180. https://0-doi-org.brum.beds.ac.uk/10.3390/catal10101180

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop