Next Article in Journal
New Synthetic Isoxazole Derivatives Acting as Potent Inducers of Fetal Hemoglobin in Erythroid Precursor Cells Isolated from β-Thalassemic Patients
Next Article in Special Issue
Efficient Catalytic Degradation of Methyl Orange by Various ZnO-Doped Lignin-Based Carbons
Previous Article in Journal
Recent Investigations on the Use of Copper Complexes as Molecular Materials for Dye-Sensitized Solar Cells
Previous Article in Special Issue
Preparation and Property Characterization of Sm2EuSbO7/ZnBiSbO5 Heterojunction Photocatalyst for Photodegradation of Parathion Methyl under Visible Light Irradiation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Synthesis of CoMoO4 Nanofibers by Electrospinning as Efficient Electrocatalyst for Overall Water Splitting

Key Laboratory for Photonic and Electronic Bandgap Materials, Ministry of Education, School of Physics and Electronic Engineering, Harbin Normal University, Harbin 150025, China
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(1), 7; https://doi.org/10.3390/molecules29010007
Submission received: 2 November 2023 / Revised: 11 December 2023 / Accepted: 14 December 2023 / Published: 19 December 2023

Abstract

:
To improve the traditional energy production and consumption of resources, the acceleration of the development of a clean and green assembly line is highly important. Hydrogen is considered one of the most ideal options. The method of production of hydrogen through water splitting constitutes the most attractive research. We synthesized CoMoO4 nanofibers by electrospinning along with post-heat treatment at different temperatures. CoMoO4 nanofibers show a superior activity for hydrogen evolution reaction (HER) and only demand an overpotential of 80 mV to achieve a current density of 10 mA cm–2. In particular, the CoMoO4 catalyst also delivers excellent performances of oxygen evolution reaction (OER) in 1 M KOH, which is a more complicated process that needs extra energy to launch. The CoMoO4 nanofibers also showed a superior stability in multiple CV cycles and maintained a catalytic activity for up to 80 h through chronopotentiometry tests. This is attributed mainly to a synergistic interaction between the different metallic elements that caused the activity of CoMoO4 beyond single oxides. This approach proved that bimetallic oxides are promising for energy production.

1. Introduction

Addressing climate change and reducing carbon emissions is conducive to promoting the green transformation of the economic structure, accelerating the production and development of green energy, mitigating the adverse effects of climate change, and reducing the losses caused to the economy and society [1,2,3]. As the foundation of hydrogen energy, electrocatalytic water splitting, which is involved in its production and utilization, has naturally become the focus of attention. For cathodic hydrogen evolution reactions, the notable material (such as Pt-based ones) catalysts are confirmed to have an excellent performance [4,5]. On the other hand, the huge potential created by the OER is what affects the production capacity. Noble-metal-based materials (Ru/IrO2) still show the highest catalytic activity toward the generation of oxygen [6,7,8]. However, the huge consumption of notable raw materials was blocking the splitting of water. Because of their abundance and excellent performance, transition metal compounds (TMCs) have been intensively studied as bifunctional electrocatalysts, especially co-based materials, such as Co-NRCNTs, Co-NCNT, CoP, Co2P, CoP/CNT, Co@N-C, Ni0.33Co0.67S2 nanowire, CoOx@CN, CoP/rGO-400, CoO/MoOx, etc. [9,10,11,12,13,14,15,16,17,18].
Different electrodes influence the reaction of water splitting, and using an alkaline solution as a condition to evolve energy is the most commercialized and feasible strategy. During the past years, transition-metal-based catalysts, such as nickel, cobalt, iron, copper, and molybdenum, have been shown to possess tunable electronic, morphological, adsorption, and structural properties, making them promising substitutes for noble metal catalysts [19,20]. In particular, metal molybdate compounds have significant desirable properties, such as non-toxicity, a low cost, and good electrochemical activity, and they have been used in some fields [21,22,23]. As shown in past reports, processed CoMoO4 obtained by the CoO/MoOx catalyst only needs a low overpotential at a current density of 10 mA cm−2. As bifunctional catalysis is important in water splitting, CoMoO4 is also considered one of the choices, due to the above merits [24,25]. From a commercialization point of view, developing a cost-efficient strategy synthesis catalyst is imperative [26,27]. Electrospinning is a controllable method and involves the process of forming jets of polymers and dissolvable materials under the action of an electric field and spinning. After heat treatment, the polymer template is sacrificed to obtain a material with a uniform fiber structure; the ordering of the crosslink structure route for charge and gas transportation can enhance the efficiency of water splitting [28,29,30,31].
Based on the above consideration, we synthesized CoMoO 4 nanofibers by electrospinning along with post-heat treatment. The CMO-650 catalyst has a stable performance through long electrochemical tests, showing a low overpotential of 80 mV to achieve a current density of 10 mA cm–2 in HER. Significantly, the catalysis performance of the CMO-650 nanofibers needs an overpotential of 370 mV at 50 mA cm−2, which is smaller than other simple oxides. Furthermore, it applies an excellent stability at a current density of 50 mA cm−2 on the alkaline electrode for 80 h. This is mainly due to the synergy between the different metal elements, resulting in the activity of CMO exceeding that of single oxides. This work proves that bimetallic oxides show promise in water splitting.

2. Results and Discussion

Figure 1 shows the X-ray diffraction (XRD) patterns of the prepared catalysts. The special peaks of CMO-650 can be attributed to the (−201), (021), (002), (−311), (−131), (−222), (400), and (040) phases of CoMoO4 (JCPDS No. 21-0868), respectively. The other samples (CMO-550, CMO-750) showed an XRD pattern similar to that of CoMoO4 with no peaks. There is a clear tendency for the crystallinity of the materials to decrease, obeying the improved temperature. This phenomenon might be attributed to the grains overlapping with the melting nanofibers and aggregating, causing the unapparent fiber structure under the high treatment.
The morphology and detailed structural information on CoMoO4 nanofibers were determined by SEM and TEM (Figure 2a–f). Figure 2a–c show SEM images of CMO-550, CMO-650, and CMO-750. The nanofibers crossed to form a network, which is clearly shown to be distinct, along with boundaries, in Figure 2a,b. The diameter of CMO-650 is around 200 nm. For CMO-750, because the temperature increases in the annealing program, the surface of nanofibers was rougher and the structure of fibers showed a state involving little melting. The CMO-650 nanofibers look like they were obtained through the right condition in order to be an electrode, with the appropriate appearance. For a definitive analysis of the surface and intrinsic composition of the CMO-650, TEM analysis was performed (Figure 2d,f). The sole nanofiber showed a uniform diameter, and the size of the fiber was same as in the shown SEM. A representative high-resolution TEM image is shown in Figure 2f, with the lattice fringe with a distance of around 0.243 nm corresponding to the (400) plane of CoMoO4, which confirms the fact that the catalyst synthesized successfully.
The analysis of valence bond changes in the surface components of the material by X-ray photoelectron spectroscopy (XPS) can demonstrate the phase transition process of the materials. As shown in Figure 3a, the peaks of Co, Mo, and O exist in the spectrum for CoMoO4. In Figure 3b, the peaks of Co 2p3/2 and Co 2p1/2 have binding energies of 780.5 and 796.6 eV, which are the characteristics of Co-O species [32,33,34]. The other peaks of Co 2p at 786.8 and 802.8 eV are shake-up satellites. As Figure 3c shows, the binding energies of 232 and 235.2 eV are matched to Mo 3d5/2 and Mo 3d3/2, which confirms the presence of the Mo VI oxidation state, which is consistent with MoO42−. As Figure 3d shows, two peaks with binding energies of 530.3 eV and 531.9 eV are matched to metal-oxygen bonds and OH- groups for O 1s [30,35,36]. These standard characteristic peaks with the right binding energies are deeply obvious, proving that the sample was successfully produced. We also tested the XPS survey spectrum of CMO-550 and CMO-750, and we found that the binding energies of Co and Mo do not have differences. We think that the calcination temperature did not impact the bonding of elements and the values of metals in the test results.
The performance of the generated hydrogen of CMO-650 (the weight of samples on the carbon paper: 2.5 mg cm–2) was examined in 1.0 M KOH. To equally detect CMO-550, CMO-750, CMO-650 powder, Co3O4, and MoO3, all the above catalysts are calculated under the same conditions and possess the same loading on the substrate. In Figure 4a, the CMO-650 electrode reaches a current density of 10 mA cm−2 with a low overpotential of 80 mV, which is better than that of the CMO-550 (130 mV), CMO-750 (103 mV), powder (139 mV), CMO-650 powder, Co3O4 (206 mV) and   MoO 3 (390 mV). The results indicate that the Mo oxides have strong synergic effects derived from the introduction of Co elements. The correlation Tafel values are calculated to be 183.43, 128.53, 172.37, 157.49, 254.69, and 300 mV dec−1 for CMO-550, CMO-650, CMO-750, CMO-650 powder, Co3O4, and MoO3, respectively (Figure 4b), which means that only a tiny overpotential change is required to meet the rapid increase in current density of CMO-650. Furthermore, electrochemical impedance spectroscopy (EIS) in Figure 4c shows that CMO-650 has the smallest charge transfer resistance (Rct), 17.67 Ω, which is an order of magnitude smaller than CMO-550 (58.7 Ω), CMO-750 (44.544 Ω), CMO-650 powder (113.78 Ω), Co3O4 (173.575 Ω) and MoO3 (403.582 Ω), indicating a fast Faradic process due to the presence of the CMO-650 interface. As shown in Figure 4d, 5000 cycles of CV are performed at a hydrogen evolution potential at a scan rate of 100 mV s−1. Testing the polarization curve of the CMO-650 after cycling shows that the potential change is almost acceptable. It is negligible and can be observed by the chronopotentiometry test by applying a current density of 50 mA cm−2 on the electrode for 80 h. CMO-650 exhibits an excellent electrochemical stability when tested for stability in alkaline environments. Figure 4e shows the different CV curves of CMO-650 at scan rates ranging from 20 to 100 mV s–1, respectively. The corresponding electric double-layer capacitance (Cdl) value is estimated by linearly fitting the change of current density with the above graph of the corresponding sweep speed. It can be noted from Figure 4f that the best Cdl value of CMO-650 is 81 mF cm–2. The CMO-550 is 54.9 mF cm–2, the CMO-750 is 25.94 mF cm–2, the powder is 60.4 mF cm–2, the Co 3 O 4 is 28.41 mF cm–2, and the MoO 3 is 2.92 mF cm–2. This result demonstrates that CMO-650 has a much higher surface area than others, which improves the efficiency of HER.
The CMO-650 catalyst displays an extraordinary performance in the HER test, but the sluggish OER kinetics still hinder commercial applications. Hence, the OER was replaced with the potential range but with the alkaline electrolyte. As Figure 5a shows, the CMO-650 electrode shows a low overpotential at a current density of 50 mA cm−2 (370 mV), which is smaller than that of the CMO-550 (410 mV), CMO-750 (416 mV), CMO-650 powder (390 mV), Co3O4 (423 mV) at 50 mA cm−2, and MoO3 (668 mV) at 25 mA cm−2. CMO-650 reaches a high current density and only needs a small overpotential, which proves that the bimetallic catalyst is better than single oxides. The Tafel slopes are calculated from the correlative LSV results, which are 78.55, 53, 78.53, 75.31, 105.77 and 170.42 mV dec−1 for CMO-550, CMO-650, CMO-750 and CMO-650 powder, respectively (Figure 5b). The smallest value of CMO-650 means that the catalyst only needs low energy to offer the current changes. The electrochemical impedance spectroscopy (EIS) renders Nyquist plots, which are fitted with a Randles circuit (Figure 5c). Herein, CMO-650 exhibited the smallest charge transfer resistance (Rct), 12.48 Ω, which is smaller than CMO-550 (21.79 Ω), CMO-750 (25.0214 Ω), CMO-650 powder (25.012 Ω), Co3O4 (150.14 Ω), and MoO3 (184.606 Ω), indicating a fast Faradic process due to the presence of the CMO-650 interface. To separate the surface area effects from the intrinsic activity, Figure 5d shows the ECSA-normalized LSV curves of materials. Interestingly, CMO nanofibers exhibited a higher intrinsic activity than the single oxides. As shown in Figure 5e, the stability of CMO-650 was tested by 8000 CV cycling, and the after-reaction LSV curve was very close to the initial curve, which shows the strong stability of CMO-650. As the chronopotentiometry test of CMO-650 shows, it can be seen that the overpotentials at a current density of 50 mA cm−2 do not show significant attenuation after 80 h (Figure 5f).
As shown in Figure 6, CoMoO4 remained on both sides of the reaction, which reveals the stability of the catalyst. More interestingly, there were no significant changes in the samples during the HER test, while the CoOOH substance (PDF 26-0480) was synthesized in the Co active site during the OER test. These results indicate that the real active materials resulting from the reconstruction of the bimetallic oxide catalyst and the coupling synergies between different metal elements enhance the electrochemical activity of the catalyst.

3. Experimental Section

3.1. Materials

Cobalt acetate tetrahydrate (C4H6CoO4·4H2O), molybdenylacetylacetonate (C10H14MoO6), potassium hydroxide (KOH), and N-dimethylformamide (DMF) were purchased from Zhiyuan Reagent Corporation (Tianjin, China, Alfa Aesar, Ward Hill, MA, USA). Polyacrylonitrile (PAN) was bought from Sigma-Aldrich Corporation (St. Louis, MO, USA). All reagents were received commercially and used without further purification.

3.2. Synthesis of CoMoO4 Nanofibers

The electrospinning solution was prepared as follows: 99.63 mg C4H6CoO4·4H2O, 130.46 mg C10H14MoO6, and a certain amount of PAN was filled in 5 g DMF. The solution was mixed for almost 24 h until it formed viscous and clear state. We replaced the solution into the electrospin device with the voltage set at 7 kV. After a day, the production was put into muffle, and we calcined the sample at 650 °C for 2 h; the purple nanofibers were named CMO-650. To convert the above sample to the corresponding CMO-X, we changed the relevant temperature (550 °C and 750 °C). For comparison, the CMO-650 powder was synthesized by directly mixing and calcining the above metal salts in muffle at 650 °C.
For comparison, we synthesized Co3O4 nanofibers using the same method with CMO-650 but not using C10H14MoO6.

3.3. Characterizations

The crystal structure of the preparations was determined by X-ray diffraction (XRD) (D/max 2600, Rigaku, Tokyo, Japan). The morphology of the materials was described with a scanning electron microscope (SEM, SU70, Hitachi, Tokyo, Japan). The atomic structure of the catalysts was observed with a transmission electron microscope (TEM, FEI, Tecnai TF20). The surface chemical qualities of the composites were measured by X-ray photoelectron spectroscopy (XPS, Thermofisher Scientific, Waltham, MA, USA).

3.4. Electrochemical Measurements

The electrochemical performance of the electrochemical workstation (VMP3): The three-electrode was fabricated using catalysts as the working electrode, and the carbon rod and Ag/AgCl were employed as the counter electrode and reference electrode, respectively. For the working electrode preparation, the catalyst (8.0 mg) and carbon black (acetylene black, 1.0 mg) were mixed in 160 μL of a 5 wt % PVDF solution under ultrasonication for 30 min. After that, apply the above mixture to the carbon paper (evenly coat 20 μL of mixture on carbon paper). All the catalysts were activated by a 50-fold CV test from 0–0.8 V with a scan rate of 50 mV s−1. The linear sweep voltammetry (LSV) was conducted from 0 to 0.6 V (vs. Ag/AgCl) at 5 mV s−1 in 1 M KOH (pH = 14). For HER, the potential ranged from 0 to −1 V (vs. Ag/AgCl) at 5 mV s−1.The LSV curves, corresponding Tafel slopes, chronopotentiometric tests, and cyclic voltammetry were obtained with iR compensation. We use the Single EIS frequency method of 100 kHz to auto iR compensation 85%. An electrochemical impedance spectroscopy (EIS) measurement was conducted at a frequency ranging from 100 kHz to 0.01 Hz at 0.5 V and −1.015 V (vs. Ag/Agcl) for HER and OER, respectively. To measure the electrochemical surface area (ECSA) of all the samples, the Cdl was calculated according to the cyclic voltammogram curves with different scan rates.

4. Conclusions

In summary, we used electrospun fiber felt as a precursor and selective calcination in air, and different nanofibers were constructed. The HER reaction was tested in an alkaline environment, and the CMO-650 showed a good activity and stability at a current density of 10 mA cm–2. Additionally, we found a good electrochemical performance in the OER test because the one-dimensional structure of CMO-650 can effectively combine electrolytes for a rapid mass transfer. The long OER test for 50 mA cm–2 can continue for 80 h. All the electrochemical tests with different oxides confirm that the unique fiber structure and bimetallic synergy are preferred to HER and OER. According to XRD, there is almost no change in the surface of the material after the HER reaction, but the material that has undergone water oxidation exhibits a new substance, which is the active substance that truly provides the activity. This work establishes an actionable strategy to provide one-dimensional CMO-650 materials to be used in bifunctional electrochemical catalysis.

Author Contributions

J.F.: Conceptualization, Methodology, Validation, Investigation, Resources; X.C.: Writing—Original draft, Visualization; L.L.: Conceptualization, Writing—Review and editing; M.Z.: Conceptualization, Validation, Writing—Review and editing, Supervision, Funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported in part by the National Natural Science Foundation of China (No. 52102228).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, J.; Guo, Z.; Liu, M.; Wang, Y.; Liu, H.; Wu, L.; Xue, Y.; Cai, N.; Li, H.; Yu, F. CoMoO4 Nanoparticles Decorated Ultrathin Nanoplates Constructed Porous Flower as an Electrocatalyst toward Overall Water Splitting and Zn-Air Batteries. Renew. Energy 2023, 212, 751–760. [Google Scholar] [CrossRef]
  2. Wang, R.; Li, F.; Ji, J.; Wang, F. CoMoO4 Enhanced Anodized Cobalt Oxide Nanotube as an Efficient Electrocatalyst for Hydrogen Evolution Reaction. Appl. Surf. Sci. 2022, 579, 152128. [Google Scholar] [CrossRef]
  3. Zhang, Z.; Ran, J.; Fan, E.; Zhou, S.; Chai, D.-F.; Zhang, W.; Zhao, M.; Dong, G. Mesoporous CoMoO4 Hollow Tubes Derived from POMOFs as Efficient Electrocatalyst for Overall Water Splitting. J. Alloys Compd. 2023, 968, 172169. [Google Scholar] [CrossRef]
  4. Ma, X.; Wei, B.; Yuan, M.; Li, J.; Liang, S.; Wu, Y.; Dai, D.; Xu, L. Self-Supported Phosphorus-Doped CoMoO4 Rod Bundles for Efficient Hydrogen Evolution. J. Mater. Sci. 2020, 55, 6502–6512. [Google Scholar] [CrossRef]
  5. Chamani, S.; Sadeghi, E.; Unal, U.; Peighambardoust, N.S.; Aydemir, U. Tuning Electrochemical Hydrogen-Evolution Activity of CoMoO4 through Zn Incorporation. Catalysts 2023, 13, 798. [Google Scholar] [CrossRef]
  6. González, D.; Heras-Domingo, J.; Sodupe, M.; Rodríguez-Santiago, L.; Solans-Monfort, X. Importance of the Oxyl Character on the IrO2 Surface Dependent Catalytic Activity for the Oxygen Evolution Reaction. J. Catal. 2021, 396, 192–201. [Google Scholar] [CrossRef]
  7. Wang, Z.; Xiao, B.; Lin, Z.; Shen, S.; Xu, A.; Du, Z.; Chen, Y.; Zhong, W. In-Situ Surface Decoration of RuO2 Nanoparticles by Laser Ablation for Improved Oxygen Evolution Reaction Activity in Both Acid and Alkali Solutions. J. Energy Chem. 2021, 54, 510–518. [Google Scholar] [CrossRef]
  8. Yu, L.; Wu, L.; McElhenny, B.; Song, S.; Luo, D.; Zhang, F.; Yu, Y.; Chen, S.; Ren, Z. Ultrafast Room-Temperature Synthesis of Porous S-Doped Ni/Fe (Oxy)Hydroxide Electrodes for Oxygen Evolution Catalysis in Seawater Splitting. Energy Environ. Sci. 2020, 13, 3439–3446. [Google Scholar] [CrossRef]
  9. Liu, Q.; Tian, J.; Cui, W.; Jiang, P.; Cheng, N.; Asiri, A.M.; Sun, X. Carbon Nanotubes Decorated with CoP Nanocrystals: A Highly Active Non-Noble-Metal Nanohybrid Electrocatalyst for Hydrogen Evolution. Angew. Chem. Int. Ed. 2014, 53, 6710–6714. [Google Scholar] [CrossRef]
  10. Wang, J.; Gao, D.; Wang, G.; Miao, S.; Wu, H.; Li, J.; Bao, X. Cobalt Nanoparticles Encapsulated in Nitrogen-Doped Carbon as a Bifunctional Catalyst for Water Electrolysis. J. Mater. Chem. A 2014, 2, 20067–20074. [Google Scholar] [CrossRef]
  11. Huang, Z.; Chen, Z.; Chen, Z.; Lv, C.; Humphrey, M.G.; Zhang, C. Cobalt Phosphide Nanorods as an Efficient Electrocatalyst for the Hydrogen Evolution Reaction. Nano Energy 2014, 9, 373–382. [Google Scholar] [CrossRef]
  12. Zou, X.; Huang, X.; Goswami, A.; Silva, R.; Sathe, B.R.; Mikmeková, E.; Asefa, T. Cobalt-Embedded Nitrogen-Rich Carbon Nanotubes Efficiently Catalyze Hydrogen Evolution Reaction at All pH Values. Angew. Chem. Int. Ed. 2014, 53, 4372–4376. [Google Scholar] [CrossRef] [PubMed]
  13. Yan, X.; Tian, L.; Atkins, S.; Liu, Y.; Murowchick, J.; Chen, X. Converting CoMoO4 into CoO/MoOx for Overall Water Splitting by Hydrogenation. ACS Sustain. Chem. Eng. 2016, 4, 3743–3749. [Google Scholar] [CrossRef]
  14. Peng, Z.; Jia, D.; Al-Enizi, A.M.; Elzatahry, A.A.; Zheng, G. From Water Oxidation to Reduction: Homologous Ni–Co Based Nanowires as Complementary Water Splitting Electrocatalysts. Adv. Energy Mater. 2015, 5, 1402031. [Google Scholar] [CrossRef]
  15. Jin, H.; Wang, J.; Su, D.; Wei, Z.; Pang, Z.; Wang, Y. In Situ Cobalt–Cobalt Oxide/N-Doped Carbon Hybrids as Superior Bifunctional Electrocatalysts for Hydrogen and Oxygen Evolution. J. Am. Chem. Soc. 2015, 137, 2688–2694. [Google Scholar] [CrossRef] [PubMed]
  16. Jiao, L.; Zhou, Y.-X.; Jiang, H.-L. Metal–Organic Framework-Based CoP/Reduced Graphene Oxide: High-Performance Bifunctional Electrocatalyst for Overall Water Splitting. Chem. Sci. 2016, 7, 1690–1695. [Google Scholar] [CrossRef] [PubMed]
  17. Wang, J.; Cui, W.; Liu, Q.; Xing, Z.; Asiri, A.M.; Sun, X. Recent Progress in Cobalt-Based Heterogeneous Catalysts for Electrochemical Water Splitting. Adv. Mater. 2016, 28, 215–230. [Google Scholar] [CrossRef]
  18. Tian, J.; Liu, Q.; Asiri, A.M.; Sun, X. Self-Supported Nanoporous Cobalt Phosphide Nanowire Arrays: An Efficient 3D Hydrogen-Evolving Cathode over the Wide Range of pH 0–14. J. Am. Chem. Soc. 2014, 136, 7587–7590. [Google Scholar] [CrossRef]
  19. Liu, Z.; Amin, H.M.A.; Peng, Y.; Corva, M.; Pentcheva, R.; Tschulik, K. Facet-Dependent Intrinsic Activity of Single Co3O4 Nanoparticles for Oxygen Evolution Reaction. Adv. Funct. Mater. 2022, 33, 2210945. [Google Scholar] [CrossRef]
  20. Soltani, M.; Amin, H.M.A.; Cebe, A.; Ayata, S.; Baltruschat, H. Metal-Supported Perovskite as an Efficient Bifunctional Electrocatalyst for Oxygen Reduction and Evolution: Substrate Effect. J. Electrochem. Soc. 2021, 168, 034504. [Google Scholar] [CrossRef]
  21. Barik, S.; Kharabe, G.P.; Illathvalappil, R.; Singh, C.P.; Kanheerampockil, F.; Walko, P.S.; Bhat, S.K.; Devi, R.N.; Vinod, C.P.; Krishnamurty, S.; et al. Active Site Engineering and Theoretical Aspects of “Superhydrophilic” Nanostructure Array Enabling Efficient Overall Water Electrolysis. Small 2023, 19, 2304143. [Google Scholar] [CrossRef]
  22. Xie, W.; Yu, T.; Ou, Z.; Zhang, J.; Li, R.; Song, S.; Wang, Y. Self-Supporting Clusters Constituted of Nitrogen-Doped CoMoO4 Nanosheets for Efficiently Catalyzing the Hydrogen Evolution Reaction in Alkaline Media. ACS Sustain. Chem. Eng. 2020, 8, 9070–9078. [Google Scholar] [CrossRef]
  23. Wang, J.; Xuan, H.; Meng, L.; Yang, J.; Yang, J.; Liang, X.; Li, Y.; Han, P. Facile Synthesis of N, S Co-Doped CoMoO4 Nanosheets as High-Efficiency Electrocatalysts for Hydrogen Evolution Reaction. Ionics 2022, 28, 4685–4695. [Google Scholar] [CrossRef]
  24. Guo, Y.; Liu, X.; Li, Y.; Ma, F.; Zhang, Q.; Wang, Z.; Liu, Y.; Zheng, Z.; Cheng, H.; Huang, B.; et al. Anion-Modulation in CoMoO4 Electrocatalyst for Urea-Assisted Energy-Saving Hydrogen Production. Int. J. Hydrogen Energy 2022, 47, 33167–33176. [Google Scholar] [CrossRef]
  25. Zang, M.; Xu, N.; Cao, G.; Chen, Z.; Cui, J.; Gan, L.; Dai, H.; Yang, X.; Wang, P. Cobalt Molybdenum Oxide Derived High-Performance Electrocatalyst for the Hydrogen Evolution Reaction. ACS Catal. 2018, 8, 5062–5069. [Google Scholar] [CrossRef]
  26. Feng, D.; Zhang, S.; Tong, Y.; Dong, X. Dual-Anions Engineering of Bimetallic Oxides as Highly Active Electrocatalyst for Boosted Overall Water Splitting. J. Colloid Interface Sci. 2022, 623, 467–475. [Google Scholar] [CrossRef] [PubMed]
  27. Yuan, C.; Cheng, G.; Ruan, W.; Ma, B.; Yuan, X.; Zhang, X.; Li, Z.; Teng, Y.; Wang, L.; Teng, F. Efficient Hydrogen Production in an Innovative S-Doped CoMoO4-Based Electrolytic Cell: 12.97% Less Energy Consumption. Sustain. Mater. Technol. 2023, 37, e00665. [Google Scholar] [CrossRef]
  28. Xu, J.; Gu, S.; Fan, L.; Xu, P.; Lu, B. Electrospun Lotus Root-like CoMoO4@Graphene Nanofibers as High-Performance Anode for Lithium Ion Batteries. Electrochim. Acta 2016, 196, 125–130. [Google Scholar] [CrossRef]
  29. Xie, S.; Wang, H.; Yao, T.; Wang, J.; Wang, C.; Shi, J.-W.; Han, X.; Liu, T.; Cheng, Y. Embedding CoMoO4 Nanoparticles into Porous Electrospun Carbon Nanofibers towards Superior Lithium Storage Performance. J. Colloid Interface Sci. 2019, 553, 320–327. [Google Scholar] [CrossRef]
  30. Li, F.; Xiao, F.; Yao, T.; Zhu, L.; Liu, T.; Lu, H.; Qian, R.; Liu, Y.; Han, X.; Wang, H. Selenizing CoMoO4 Nanoparticles within Electrospun Carbon Nanofibers towards Enhanced Sodium Storage Performance. J. Colloid Interface Sci. 2021, 586, 663–672. [Google Scholar] [CrossRef]
  31. Chang, L.; Chen, S.; Fei, Y.; Stacchiola, D.J.; Hu, Y.H. Superstructured NiMoO4 @CoMoO4 Core-Shell Nanofibers for Supercapacitors with Ultrahigh Areal Capacitance. Proc. Natl. Acad. Sci. USA 2023, 120, e2219950120. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, F.; Zhao, J.; Tian, W.; Hu, Z.; Lv, X.; Zhang, H.; Yue, H.; Zhang, Y.; Ji, J.; Jiang, W. Morphology-Controlled Synthesis of CoMoO4 Nanoarchitectures Anchored on Carbon Cloth for High-Efficiency Oxygen Oxidation Reaction. RSC Adv. 2019, 9, 1562–1569. [Google Scholar] [CrossRef] [PubMed]
  33. Zhao, J.; Ren, X.; Ma, H.; Sun, X.; Zhang, Y.; Yan, T.; Wei, Q.; Wu, D. Synthesis of Self-Supported Amorphous CoMoO4 Nanowire Array for Highly Efficient Hydrogen Evolution Reaction. ACS Sustain. Chem. Eng. 2017, 5, 10093–10098. [Google Scholar] [CrossRef]
  34. Amin, H.M.A.; Bondue, C.J.; Eswara, S.; Kaiser, U.; Baltruschat, H. A Carbon-Free Ag–Co3O4 Composite as a Bifunctional Catalyst for Oxygen Reduction and Evolution: Spectroscopic, Microscopic and Electrochemical Characterization. Electrocatalysis 2017, 8, 540–553. [Google Scholar] [CrossRef]
  35. Zhao, S.; Berry-Gair, J.; Li, W.; Guan, G.; Yang, M.; Li, J.; Lai, F.; Corà, F.; Holt, K.; Brett, D.J.L.; et al. Hydrogen Evolution: The Role of Phosphate Group in Doped Cobalt Molybdate: Improved Electrocatalytic Hydrogen Evolution Performance (Adv. Sci. 12/2020). Adv. Sci. 2020, 7, 2070067. [Google Scholar] [CrossRef]
  36. Geng, S.; Chen, L.; Chen, H.; Wang, Y.; Ding, Z.-B.; Cai, D.; Song, S. Revealing the Electrocatalytic Mechanism of Layered Crystalline CoMoO4 for Water Splitting: A Theoretical Study from Facet Selecting to Active Site Engineering. Chin. J. Catal. 2023, 50, 334–342. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of CMO-550, CMO-650, and CMO-750.
Figure 1. XRD patterns of CMO-550, CMO-650, and CMO-750.
Molecules 29 00007 g001
Figure 2. SEM images of (a) CMO-550, (b) CMO-650, and (c) CMO-750. (d,e) TEM images of CMO-650; (f) HR-TEM lattice image of CMO-650.
Figure 2. SEM images of (a) CMO-550, (b) CMO-650, and (c) CMO-750. (d,e) TEM images of CMO-650; (f) HR-TEM lattice image of CMO-650.
Molecules 29 00007 g002
Figure 3. (a) XPS survey spectrum for CMO-650. XPS spectra of CMO-650 in the (b) Co 2p, (c) Mo 3d, and (d) O 1s.
Figure 3. (a) XPS survey spectrum for CMO-650. XPS spectra of CMO-650 in the (b) Co 2p, (c) Mo 3d, and (d) O 1s.
Molecules 29 00007 g003
Figure 4. (a) LSV curves for CMO-550, CMO-750, CMO-650 powder, Co3O4, and MoO3 for HER with a scan speed of 5 mV s−1. (b) Tafel slopes for CMO-550, CMO-750, CMO-650 powder, Co3O4, and MoO3. (c) EIS pattern of the above catalysts. (d) The stability of CMO-650 initially and after 5000 cycles with the inline diagram of the chronopotentiometry test with a current density of 50 mA cm−2 for 80 h. (e) The different CV curves of CMO-650 ranging from 20–100 mV s−1. (f) The line fitter between the scan rates and current densities of the above samples.
Figure 4. (a) LSV curves for CMO-550, CMO-750, CMO-650 powder, Co3O4, and MoO3 for HER with a scan speed of 5 mV s−1. (b) Tafel slopes for CMO-550, CMO-750, CMO-650 powder, Co3O4, and MoO3. (c) EIS pattern of the above catalysts. (d) The stability of CMO-650 initially and after 5000 cycles with the inline diagram of the chronopotentiometry test with a current density of 50 mA cm−2 for 80 h. (e) The different CV curves of CMO-650 ranging from 20–100 mV s−1. (f) The line fitter between the scan rates and current densities of the above samples.
Molecules 29 00007 g004
Figure 5. (a) LSV curves for CMO-550, CMO-750, CMO-650 powder, Co3O4, and MoO3 for OER with a scan speed of 5 mV s−1. (b) Tafel plots for CMO-550, CMO-750, CMO-650 powder, Co3O4 and   MoO 3 . (c) EIS pattern of the above catalysts. (d) The stability of CMO-650 initially and after 5000 cycles with the inline diagram of the chronopotentiometry test with a current density of 50 mA cm−2 for 80 h. (e) The different CV curves of CMO-650 ranging from 20–100 mV s−1. (f) The line fitter between the scan rates and current densities of the above samples.
Figure 5. (a) LSV curves for CMO-550, CMO-750, CMO-650 powder, Co3O4, and MoO3 for OER with a scan speed of 5 mV s−1. (b) Tafel plots for CMO-550, CMO-750, CMO-650 powder, Co3O4 and   MoO 3 . (c) EIS pattern of the above catalysts. (d) The stability of CMO-650 initially and after 5000 cycles with the inline diagram of the chronopotentiometry test with a current density of 50 mA cm−2 for 80 h. (e) The different CV curves of CMO-650 ranging from 20–100 mV s−1. (f) The line fitter between the scan rates and current densities of the above samples.
Molecules 29 00007 g005
Figure 6. XRD patterns of CMO-650 after stability test; (a) after HER and (b) after OER.
Figure 6. XRD patterns of CMO-650 after stability test; (a) after HER and (b) after OER.
Molecules 29 00007 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Fan, J.; Chang, X.; Li, L.; Zhang, M. Synthesis of CoMoO4 Nanofibers by Electrospinning as Efficient Electrocatalyst for Overall Water Splitting. Molecules 2024, 29, 7. https://doi.org/10.3390/molecules29010007

AMA Style

Fan J, Chang X, Li L, Zhang M. Synthesis of CoMoO4 Nanofibers by Electrospinning as Efficient Electrocatalyst for Overall Water Splitting. Molecules. 2024; 29(1):7. https://doi.org/10.3390/molecules29010007

Chicago/Turabian Style

Fan, Jiahui, Xin Chang, Lu Li, and Mingyi Zhang. 2024. "Synthesis of CoMoO4 Nanofibers by Electrospinning as Efficient Electrocatalyst for Overall Water Splitting" Molecules 29, no. 1: 7. https://doi.org/10.3390/molecules29010007

Article Metrics

Back to TopTop