Next Article in Journal
Synthesis, Copper(II) Binding, and Antifungal Activity of Tertiary N-Alkylamine Azole Derivatives
Previous Article in Journal
Growth, Spectroscopic Characterization and Continuous-Wave Laser Operation of Er,Yb:GdMgB5O10Crystal
Previous Article in Special Issue
Enhancing Supercapacitor Performance with Zero-Dimensional Tin–Niobium Oxide Heterostructure Composite Spheres: Electrochemical Insights
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Investigation of the Interface between Transition Metal Oxides (MnOx, FeOx, CoOx and NiOx)/MoO3 Composite Electrocatalysts for Oxygen Evolution Reactions

School of Chemical Engineering, Yeungnam University, Gyeongsan 38541, Republic of Korea
*
Authors to whom correspondence should be addressed.
Inorganics 2024, 12(9), 241; https://doi.org/10.3390/inorganics12090241
Submission received: 1 August 2024 / Revised: 24 August 2024 / Accepted: 27 August 2024 / Published: 2 September 2024
(This article belongs to the Special Issue Mixed Metal Oxides II)

Abstract

:
This study presents the synthesis of a multicomponent metal oxide electrocatalyst that increases the activity of the oxygen evolution reaction (OER). We synthesized transition metal oxides (MnOx, FeOx, CoOx, and NiOx) with MoO3 heterostructures through a solid-state reaction approach at low cost. In comparison to the other compositions, CoOx garnered higher attention and demonstrated superior performance on account of its large surface area and varied crystal facets. The MnOx-MoO3, FeOx-MoO3, CoOx-MoO3, and NiOx-MoO3 compositions attained an overpotential of 390 mV, 350 mV, 310 mV, and 340 mV, respectively, at a current density of 10 mA cm−2 in alkaline solution. The performance of OER was enhanced in CoOx-MoO3 at 10 mA cm−2, while FeOx-MoO3 exhibited a lower current density at 100 mA cm−2 than other metal oxides. The CoOx-MoO3 material exhibited a favorable crystal interface transition due to the presence of MoO3 oxide. For the first time, we report on the MoO3-to-(MnOx, FeOx, CoOx, and NiOx) interface crystal transition and the active surface area for OER catalytic activity in water-splitting processes. This investigation intends to develop an electrocatalyst that is capable of producing hydrogen with the use of heterostructure metal oxides.

Graphical Abstract

1. Introduction

Electrochemical energy storage provides a sustainable solution for generating green energy without harmful carbon emissions. One of the most significant non-carbon emission processes is the water-splitting process [1,2]. This process involves two main reactions, namely the hydrogen evolution reaction (HER) and the oxygen evolution reaction (OER), which produce hydrogen and oxygen, respectively, through the decomposition of water [3,4]. The OER reaction requires four electrons and has a high overpotential, making it a challenging issue in water decomposition. Therefore, many researchers have focused on developing OER electrocatalysts for electrochemical reactions. RuO2 and IrO2 are used in commercial applications [5]. However, these benchmark catalysts are lacking in resources and are costly. Therefore, electrocatalysts based on various transition metal composites, such as oxides, sulfides, nitrides, phosphides, hydroxides, and carbides, have been reported as potential replacements for benchmark catalysts [6,7]. These composite electrocatalysts, composed of transition metals, are facing challenges related to surface area, stability, and electrical conductivity. Therefore, research is needed to develop more innovative electrocatalysts in the transition metal composite by tuning composition, morphology, and size. The high surface area and good electrical conductivity of electrocatalysts boost their catalytic activity significantly [8]. Multi-metal oxide composites enhance surface area and electrical conductivity, boosting OER activity. This is because the OER reaction takes place on the surface of the electrode materials. The interface electrode materials play a significant role in improving electrokinetic reactions [9]. Recent reports indicate that molybdenum oxide (MoO3), a key component in electrocatalytic systems, can exchange surface oxygen ions reversibly in a heterostructure interface. This property can aid in charge transfer and the production of O* during the OER process. MoO3 has a higher number of active sites and easy charge transport during electrocatalysis compared to platinum. The RuO2/MoO3 composite heterostructure exhibited a low overpotential and 3.2 times greater activity [10]. Yi Liu et al. researched the integration of MoO3 patches into Ni oxyhydroxide nanosheets [11]. At 10 mA cm−2, this composite has an overpotential of 260 mV. According to the findings, the Ni-Mo heterostructure increases electrocatalytic activity. Yao Zhang et al. studied the multicomponent interface structure (Fe-S-NiMoO4/MoO3@NF) for OER activity. The catalyst worked at an overpotential of 271 at 500 mA cm−2. In this composite, the electric conductivity improved the MoO3 structure by doping Fe [12]. Li et al. developed a bifunctional catalyst of MoO3-Ni-NiO via the electrodeposition method. It showed an OER overpotential of 347 mV at 100 mA cm−2. The MoO3-Ni-NiO interface composite revealed a low energetic barrier and boosted the catalytic activity [13]. Tariq et al. investigated low noble metal materials with the addition of MoO3. The IrO2/MoO3 interface composite showed long-term stability up to 40,000 s in acidic conditions. The composite resulted in a two-fold enhancement at the current density and a seven-fold increase in mass activity [14]. The above studies indicate that MoO3 exhibits good interface activity with other metal oxides. However, the heterostructure functions and electrocatalytic activity of MoO3 with transition metal oxides (Mn, Fe, Co, and Ni) are not fully understood.
Therefore, a multicomponent interface heterostructure was investigated in this research paper via a simple solid-state reaction method. The heterostructure was constructed using transition metal oxides (Mn, Fe, Co, and Ni) integrated with MoO3. The performance of MnOx-MoO3, FeOx-MoO3, CoOx-MoO3, and NiOx-MoO3 for OER was analyzed and compared. The CoOx-MoO3 heterostructure exhibited a higher surface area (57.05 m2 cm−1) than the other composite electrocatalysts. The analysis of OER revealed that the CoOx-MoO3 composite exhibited a low overpotential (310 mV) and was more active than the other tested heterostructures at 10 mA cm−2. However, the FeOx-MoO3 composite interface showed higher activity at 100 mA cm−2 than cobalt oxide. Based on the results of this study, an electrocatalyst comprising MoO3 coupled with CoOx and FeOx composites leads to a more significant interface transition for OER activity than other catalysts. The hybrid composite electrocatalyst of MoO3 with a transition metal oxide enhances OER performance. This rational design of the heterostructure interface improves alkaline stable electrocatalysts for OER in water-splitting reactions.

2. Materials and Methods

2.1. Materials

The materials included nickel (II)acetate tetrahydrate (97%, Daejung, Siheung, Republic of Korea), Ferrous acetate, Cobalt acetate, manganese nitrate, ammonium molybdate tetrahydrate (97%, Daejung, Republic of Korea), carbon black–acetylene black 100% compressed (CA), specific surface area 75 m2/g (Alfa Aesar, Seoul, Republic of Korea), polyvinylidene fluoride (PVDF, Mw ~534,000, Sigma-Aldrich, Seoul, Republic of Korea), 1-methly 2 pyrrolidone (NMP, 99%, extra pure, Duksan chemicals, Ansan-si, Republic of Korea), potassium hydroxide (85%, KOH, Duksan chemicals, Republic of Korea), demineralized water (DI). All the chemicals used in this experiment were received as they were, without any additional purification process.

2.2. Synthesis of Active Metal Oxide Composites

The metal precursors were mixed with ammonium molybdate tetrahydrate with an equal molar ratio. The mixture was ground using a mortar and pestle for 20 min. Afterward, carbon (0.1 g) was added to the above reactant for further grinding and mixed well. Finally, the reactant was transferred into a ceramic crucible for thermal treatment. The crucible was placed in the air oven at 300 °C for 4 h with a temperature increase of 2 °C per minute. After the completion of the reaction, the product was ground for 20 min using a mortar and pestle again. Likewise, each metal oxide composite (MnOx-MoO3, FeOx-MoO3, CoOx-MoO3, and NiOx-MoO3) was synthesized separately.

2.3. Physiochemical Characterization

BET (Brunauer–Emmett–Teller) surface areas (m2/g) were determined through nitrogen adsorption–desorption analysis. The nature of crystal structure type and composition of metal oxide was investigated using X-ray diffraction (XRD) with an Xpert Pro device. The device was equipped with a Cu Kα radiation source with a wavelength of 1.5406 Å and a step size of 0.02°. The oxidation state of as-prepared materials was analyzed by X-ray photoelectron spectroscopy (XPS) using Thermo Scientific (Waltham, MA, USA) K-α surface analysis. To analyze high-resolution transmission electron microscopic (HRTEM) images, we utilized a transmission electron microscope with JEM-2100 (Tokyo, Japan). For distinguishing different particles, high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) and energy dispersive X-ray (EDX, Bruker Nano GmbH, Berlin, Germany) element mapping images were used. Carbon hybridization was confirmed by the G and D bands in Raman spectroscopy using an XploRA plus device, HORIBA Jobin Yuon S.A.S, France. The scanning electron microscopy (SEM) analysis was conducted using a HITACHI S-4800, Tokyo, Japan, instrument equipped with an energy-dispersive X-ray spectroscopy (EDAX) system to examine the surface microstructures and particle size.

2.4. Preparation of Working Electrodes

The electrode was prepared using carbon cloth (CC). The CC was cleaned by thermal treatment at 400 °C. Next, it was immersed in ethanol, acetone, and DI water for 20 min each. Finally, the CC was dried overnight at 60 °C. The electrocatalyst slurry, consisting of 90% active catalyst, 10% PVDF, and NMP solvent, was applied to a pretreated CC substrate measuring 1 cm × 1 cm. The slurry was then dried overnight at 60 °C.

2.5. Electrochemical Analysis

All electrochemical measurements were conducted at room temperature in a 1 M KOH aqueous solution using a three-electrode system. The system comprised a working electrode, a Pt sheet as the counter electrode, and a Hg/HgO electrode as the reference electrode. The measurements were carried out using a Corretest electrochemical workstation. Electrochemical impedance spectroscopy (EIS) was studied at an open-circuit potential at frequencies ranging from 1 Hz to 100 kHz. Chronopotentiometry curves were recorded at a constant potential without iR compensation. Cyclic voltammetry (CV) analyses were carried out at different scan rates (5, 10, 15, 20, and 25 mV s−1) at room temperature. The electrochemical double-layer capacitance (Cdl) and electrochemical surface area (ECSA) were calculated from different scan rates of CV [15] as follows:
E C S A = C d l C s
where Cs is a specific capacitance of 1 M KOH solution (0.040 mF cm−2).
Linear sweep voltammetry (LSV) was recorded at 5 mV/S scan rate. All recorded potentials were converted into reversible hydrogen electrodes (RHEs) using Equation (1).
E R H E = E H g / H g O + 0.0596   X   P H + E H g / H g O 0

3. Results

XRD analysis and Raman spectroscopy were used to establish the crystal structure and content. The MoO3 diffraction line in Figure 1a matches the as-synthesized MnOx-MoO3, FeOx-MoO3, CoOx-MoO3, and NiOx-MoO3. The diffraction angles indexed as 12.7 (001), 23.3 (100), 25.7 (002), and 27.7 (011) on JCPDS Card Nos. 85-2407 identify the MoO3 crystal planes [14]. The diffraction pattern verifies that the MoO3 crystal plane overlaps with that of the composite metal oxide. The diffractogram does not reveal the diffraction peaks for the nanostructures of MnOx, FeOx, CoOx, and NiOx because of overlaps with the MoO3 peak and small concentration, as also confirmed by XPS [16]. The intensity of MoO3 diffraction peaks is decreased in the composite metal oxide diffraction pattern, which is determined by the metal oxide composite combined with MoO3. However, HRTEM, XPS, and SAED analyses confirm the presence of this metal oxide composite. The carbon substance’s dislocation was not detected by XRD. Carbon and chemical alterations in MoO3 samples were examined using Raman spectroscopy, as shown in Figure 1b. The carbon structural properties and dislocations were largely determined by the maximum intensity and position matching. The D and G band intensity ratios of the samples were calculated at 1.2, 1.0, 1.1, and 0.91 for MnOx-MoO3, FeOx-MoO3, CoOx-MoO3, and NiOx-MoO3, respectively [17,18]. These intensity ratios indicate that the carbon surface was delocalized by metal oxide. Mn, Co, and Ni metal oxides have a strong affinity with the edge group of carbon. However, the Fe metal oxide’s affinity with carbon was very low. Figure 1c shows N2 adsorption–desorption isotherms for specific surface area (SSA) analysis. It was found that the SSA of CoOx-MoO3 (57.05 m2 g−1) was higher than that of MnOx-MoO3 (37.42 m2 g−1), FeOx-MoO3 (47.04 m2 g−1), and NiOx-MoO3 (45.44 m2 g−1).
The porosity of the active catalyst could determine the surface area of the metal oxide composition. As a result, the high surface area obtained in CoOx-MoO3 was due to the multi-crystal facet morphology, which was confirmed in the TEM image [19]. Furthermore, we calculated the sample’s pore size using the Barrett–Joyner–Halenda (BJH) plots, which were created using the data obtained from the BET surface area technique [20,21]. Figure 1d presents the mesopore in the range of 12 to 14 nm for all samples (Table S1). Compared to other samples, the higher surface area of CoOx-MoO3 led to a higher pore volume (0.1805 cm3) and pore diameter (14.9 nm). This pore diameter facilitated the formation of an activation site during an electrocatalytic reaction. CoOx-MoO3 had a higher SSA and cumulative volume in the mesopores than other metal oxide composites [22].
The XPS analysis was used to characterize the surface chemical composition of the as-synthesized materials. The Mn 2p, Fe 2p, Co 2p, and Ni 2p spectra of MnOx-MoO3, FeOx-MoO3, CoOx-MoO3, and NiOx-MoO3, respectively, reveal the elemental composition of the metal species present in the composite. Figure 2a displays two peaks located at the binding energy of 653.37 and 641.63 eV, which correspond to Mn 2p1/2 and Mn 2p3/2, respectively. The presence of Mn3+ was confirmed with a binding energy difference of 11.7 eV between Mn 2p1/2 and Mn 2p3/2 [23]. However, satellite peaks at 646.5 and 657.3 eV for Mn 2p3/2 and Mn 2p1/2, respectively, are deconvoluted [24,25]. In Figure 2b, the XPS spectrum of Fe is split into two energy bands, namely Fe 2p3/2 (lower energy band) and Fe 2p1/2 (higher energy band). The peaks are observed at 711.7 eV (Fe 2p3/2) and 724.6 eV (Fe 2p1/2) in Fe3+ and Fe2+ core lines, respectively [26,27]. Additionally, the peaks located at approximately 718.9 and 714.5 eV are the satellite peaks associated with them [23,28]. Figure 2c illustrates that Co 2p is deconvoluted into two major peaks, and two small satellite peaks are observed. The two major peaks are located at 780.6 eV and 796.2, corresponding to the Co3+ core line in Co 2p3/2 and Co 2p1/2, respectively. The satellite peaks are observed at ~785.1 eV (Co 2p3/2) and ~803.6 eV (Co 2p1/2), which corresponds to the core line of Co2+ [29,30]. Figure 2d shows two component peaks in Ni 2p spectra. The first main peak matches with Ni 2p3/2 and is located at 854.4 eV, whereas the second main peak matches with Ni 2p1/2 spin-orbit levels of NiO and is centered at 871.6 eV. The Ni2+ state is represented by the peaks that match with Ni 2p3/2 [30]. Furthermore, at 880.7, 877.7, 867.6, and 863.9 eV, the characteristic satellite peaks are detected as deconvoluted [18,31].
The composite structure of MnOx-MoO3 and FeOx-MoO3 in Figure 3 was confirmed by the HRTEM image and elemental mapping. In Figure 3a,b, MnOx and MoO3 crystal facets are observed on the carbon surface. The C, O, Mn, and Mo elemental mapping in Figure 3c–f confirms the elemental and atomic composition present in MnOx-MoO3. Figure 3g,h show the image of FeOx-MoO3. The elemental composition was confirmed by the mapping image presented in Figure 3i–l. The crystallite metal oxides were dispersed in the amorphous carbon network. Figure 4a,b show images of CoOx-MoO3, while Figure 4g,h show images of NiOx-MoO3. In the TEM image, CoOx-MoO3 appears as a multi-faceted crystal, and its surface area is increased compared to other materials. This increase was confirmed in the BET analysis. The elemental composition of C, O, Co, and Mo is confirmed in the elemental mapping images in Figure 4c–f. Similarly, the mapping images in Figure 4i–l show the composition of NiOx-MoO3. The interplanar spacing diagram and chemical composition for MnOx-MoO3, FeOx-MoO3, CoOx-MoO3, and NiOx-MoO3 are shown in Figures S1–S4. The FE-SEM image in Figure S5 shows the spherical shape of particles in nanometers.
A study was conducted to analyze the electrochemical performance of an all-composite nanostructure and compare its performance with various other similar nanostructures. Using linear sweep voltammetry (LSV) experiments, the OER activity performance was measured in a 1 M KOH medium at a scan rate of 5 mV s−1. The performance of the all-composite nanostructure regarding OER is presented in Figure 5a. The CoOx-MoO3 LSV curve indicates that the overpotential required to obtain a current density of 10 mA cm−2 is 310 mV. This value is significantly lower than its counterparts such as MnOx-MoO3 (390 mV), FeOx-MoO3 (350 mV), and NiOx-MoO3 (340 mV). Electrochemical impedance spectroscopy (EIS) was used to gain a better understanding of the catalytic activity of all catalysts in terms of OER, as shown in Figure 5b. The CoOx-MoO3 heterostructure has a much smaller semicircular area than the MnOx-MoO3, FeOx-MoO3, and NiOx-MoO3 heterostructures. This suggests that the CoOx-MoO3 heterostructure has a lower impedance (Rct). It was predicted that the CoOx-MoO3 heterostructure would have a much smaller Rct value (0.205 Ω) than MnOx-MoO3 (0.285 Ω), FeOx-MoO3 (0.911 Ω), and NiOx-MoO3 (0.259 Ω). This means that it conducts electricity better and moves charges faster. The equivalent circuit pattern and fitting curves are given in Figure S6. As shown in Figure 5c, the OER kinetic was examined using Tafel slope analysis. The CoOx-MoO3 catalyst exhibited a Tafel slope of 146 mV dec−1, indicating superior OER activity compared to its constituents MnOx-MoO3, FeOx-MoO3, and NiOx-MoO3, which exhibited Tafel slopes of 258, 103, and 125 mV dec−1, respectively. Figure 5d shows the comparative analysis of OER overpotential with all active catalysts. Figure 6a shows the Cdl value obtained from CV analysis at different scan rates (5, 10, 15, 20, and 25 mV s−1). The ECSA calculated from the Cdl of a specific metal oxide composite is illustrated in Figure 6b. A high ECSA value enhances the electrokinetic interaction between the electrode and the electrolyte, leading to improved OER activity. Therefore, the OER activity of Co and Fe was enhanced with the MoO3 composite electrocatalyst.
The chronopotentiometry curve of the CoOx-MoO3 heterostructure was plotted at a current density of 10 mA cm−2 for 24 h of electrolysis, which is presented in Figure 7a. It shows strong stability in the alkaline-mediated electrolysis. The XRD patterns before and after the stability test of CoOx-MoO3 are illustrated in Figure 7b. The mixed cobalt oxide peaks change into Co3O4 peaks at (31.3° and 36.7°) (JCPDS No. 48–1719). The characteristic diffraction pattern of MoO3 is not observed by XRD clearly, demonstrating that the MoO3 layer is amorphous. Its peak is merged with CC peaks at 25°.

4. Discussion

The electrocatalytic activity of FeOx and CoOx with MoO3 exhibited significant OER activity due to the heterostructure of these metal oxides on a carbon substrate when compared to other metal oxides, as indicated in Table 1. As the atomic number increased in the transition metal 3d series from Mn to Ni, the high spin decreased. Together with the ligand, MnOx(d5) and MoO3(d0) produced a high-spin stable oxide composite. Because of Mn’s high-spin state, the interaction between Mn and Mo was quite weak in this instance. Consequently, the OER activity of the MnOx-MoO3 composite electrocatalyst decreased. However, the presence of a low-spin composite increased the interaction between FeOx (d6) and CoOx(d7) with MoO3. At the electrode surface, an electrokinetic channel was created by the interaction of the low-spin d orbital electrons of Fe and Co with MoO3. The intermediate performance of Co and Fe with MoO3 was observed in NiOx(d8). The electrical distribution of the metal oxide interface was altered by the addition of MoO3 to the structure, which improved OER performance [19]. The metal oxide–metal oxide interface’s electron-transfer channel, which was made possible by orbital delocalization in the composite catalyst, significantly increased reaction kinetics. The unique interfacial contact and synergistic catalytic effects were found to ensure strong intrinsic performance and higher stability for electrochemical water splitting.

5. Conclusions

A synthesized composite catalyst demonstrated electrocatalytic activity in the OER for water electrolysis. The analysis using XRD, XPS, Raman spectroscopy, and HRTEM confirmed the formation of a metal oxide composite. The CoOx-MoO3 composite exhibited a high surface area, which enhanced its electrocatalytic activity. This resulted in a 310 mV overpotential at 10 mA cm−2. The FeOx-MoO3 composite showed the best performance in OER at 100 mA cm−2 compared to all other catalysts. The effectiveness of higher OER activity is directly related to the order of CoOx-MoO3, FeOx-MoO3, NiOx-MoO3, and MnOx-MoO3. This is explained by the synergy action of the conductive carbon support and the metal oxide–MoO3 heterostructure. An increase in the number of accessible active sites, better charge transmission, and increased structural durability led to better OER performance. This study paves the way for exploring improved electrocatalyst activity in metal oxide composites via heterojunctions for OER in overall water splitting without using noble metals.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/inorganics12090241/s1, The supporting information includes a table of surface area and porosity, SAED images and EDS patterns, FESEM images, particle size distribution, and electrochemical impedance spectra of MnOx-MoO3, FeOx-MoO3, CoOx-MoO3, and NiOx-MoO3. Table S1: BET surface area, pore volume, and pore diameter of all active catalyst materials. Figure S1: (a) SAED image of MnOx-MoO3 and (b) EDS pattern of MnOx-MoO3. Figure S2: (a) SAED image of FeOx-MoO3 and (b) EDS pattern of FeOx-MoO3. Figure S3: (a) SAED image of CoOx-MoO3 and (b) EDS pattern of CoOx-MoO3. Figure S4: (a) SAED image of NiOx-MoO3 and (b) EDS pattern of NiOx-MoO3. Figure S5: FE-SEM image of (a) MnOx-MoO3, (b) FeOx-MoO3, (c) CoOx-MoO3, (d) NiOx-MoO3 of active catalysts, (e) Particle size distribution histogram. Figure S6: Electrochemical impedance spectra with fitting of all active catalysts.

Author Contributions

Conceptualization, K.D.; methodology, K.D. and G.S.; software, M.B.; data curation, M.B. and T.S; formal analysis, T.S.; resources, G.S; writing—original draft preparation, K.D.; writing—review and editing, G.S. and T.H.O.; visualization, T.S.; supervision, T.H.O.; funding acquisition, T.H.O.; project administration, T.H.O. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Research Foundation of Korea grant sponsored by the Korean government (MIST) (No. 2022R1A2C1004283) and the Korea Basic Science Institute (National Research Facilities and Equipment Center) grant funded by the Ministry of Education (No. 2019R1A6C1010046).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Epelle, E.I.; Desongu, K.S.; Obande, W.; Adeleke, A.A.; Ikubanni, P.P.; Okolie, J.A.; Gunes, B. A comprehensive review of hydrogen production and storage: A focus on the role of nanomaterials. Int. J. Hydrogen Energy 2022, 47, 20398–20431. [Google Scholar] [CrossRef]
  2. Sriram, G.; Dhanabalan, K.; Ajeya, K.V.; Aruchamy, K.; Ching, Y.C.; Oh, T.H.; Jung, H.Y.; Kurkuri, M. Recent progress in anion exchange membranes (AEMs) in water electrolysis: Synthesis, physio-chemical analysis, properties, and applications. J. Mater. Chem. A Mater. 2023, 11, 20886–21008. [Google Scholar] [CrossRef]
  3. Dhanabalan, K.; Sriram, G.; Sannasi, V.; Ajeya, K.V.; Jung, S.H.; Jung, H.Y. Water splitting catalysts using non-metal covalent organic frameworks: A review. Int. J. Hydrogen Energy 2024, 51, 376–398. [Google Scholar] [CrossRef]
  4. Dhanabalan, K.; Perumalsamy, M.; Sriram, G.; Murugan, N.; Sadhasivam, T.; Oh, T.H. Metal–Organic Framework (MOF)-Derived Catalyst for Oxygen Reduction Reaction (ORR) Applications in Fuel Cell Systems: A Review of Current Advancements and Perspectives. Energies 2023, 16, 4950. [Google Scholar] [CrossRef]
  5. Lee, Y.; Suntivich, J.; May, K.J.; Perry, E.E.; Shao-Horn, Y. Synthesis and Activities of Rutile IrO2 and RuO2 Nanoparticles for Oxygen Evolution in Acid and Alkaline Solutions. J. Phys. Chem. Lett. 2012, 3, 399–404. [Google Scholar] [CrossRef] [PubMed]
  6. Parra-Puerto, A.; Ng, K.L.; Fahy, K.; Goode, A.E.; Ryan, M.P.; Kucernak, A. Supported Transition Metal Phosphides: Activity Survey for HER, ORR, OER, and Corrosion Resistance in Acid and Alkaline Electrolytes. ACS Catal. 2019, 9, 11515–11529. [Google Scholar] [CrossRef]
  7. Zhou, Y.; Sun, S.; Song, J.; Xi, S.; Chen, B.; Du, Y.; Fisher, A.C.; Cheng, F.; Wang, X.; Zhang, H.; et al. Enlarged Co—O Covalency in Octahedral Sites Leading to Highly Efficient Spinel Oxides for Oxygen Evolution Reaction. Adv. Mater. 2018, 30, 1802912. [Google Scholar] [CrossRef] [PubMed]
  8. Wang, H.; Pei, Y.; Wang, K.; Zuo, Y.; Wei, M.; Xiong, J.; Zhang, P.; Chen, Z.; Shang, N.; Zhong, D.; et al. First-Row Transition Metals for Catalyzing Oxygen Redox. Small 2023, 19, 2304863. [Google Scholar] [CrossRef]
  9. Kumar Ni Mishra, D.; Gi Seo, S.; Na, T.; Jin, S.H. Hierarchical formation of Ni sulfide single walled carbon nanotubes heterostructure on tin-sulfide scaffolds via mediated SILAR process: Application towards long cycle-life solid-state supercapacitors. Ceram. Int. 2022, 48, 16656–16666. [Google Scholar] [CrossRef]
  10. Tariq, M.; Wu, Y.; Ma, C.; Ali, M.; Zaman, W.Q.; Abbas, Z.; Ayub, K.S.; Zhou, J.; Wang, G.; Cao, L.M.; et al. Boosted up stability and activity of oxygen vacancy enriched RuO2/MoO3 mixed oxide composite for oxygen evolution reaction. Int. J. Hydrogen Energy 2020, 45, 17287–17298. [Google Scholar] [CrossRef]
  11. Liu, Y.; Liu, P.; Men, Y.L.; Li, Y.; Peng, C.; Xi, S.; Pan, Y. XIncorporating MoO3 Patches into a Ni Oxyhydroxide Nanosheet Boosts the Electrocatalytic Oxygen Evolution Reaction. ACS Appl. Mater. Interfaces 2021, 13, 26064–26073. [Google Scholar] [CrossRef] [PubMed]
  12. Zhang, Y.; Guo, H.; Ren, J.; Li, X.; Ren, W.; Song, R. MoO3 crystal facets modulation by doping heteroatom Fe from polyoxometalate for quasi-industrial oxygen evolution reaction. Appl. Catal. B 2021, 298, 120582. [Google Scholar] [CrossRef]
  13. Li, X.; Wang, Y.; Wang, J.; Da, Y.; Zhang, J.; Li, L.; Zhong, C.; Deng, Y.; Han, X.; Hu, W. Sequential Electrodeposition of Bifunctional Catalytically Active Structures in MoO3/Ni–NiO Composite Electrocatalysts for Selective Hydrogen and Oxygen Evolution. Adv. Mater. 2020, 32, 2003414. [Google Scholar] [CrossRef]
  14. Tariq, M.; Zaman, W.Q.; Sun, W.; Zhou, Z.; Wu, Y.; Cao, L.M.; Yang, J. Unraveling the Beneficial Electrochemistry of IrO2/MoO3 Hybrid as a Highly Stable and Efficient Oxygen Evolution Reaction Catalyst. ACS Sustain. Chem. Eng. 2018, 6, 4854–4862. [Google Scholar] [CrossRef]
  15. Bhosale, M.; Thangarasu, S.; Murugan, N.; Kim, Y.A.; Oh, T.H. Engineering 2D heterostructured VS2-rGO-Ni nanointerface to stimulate electrocatalytic water splitting and supercapacitor applications. J. Energy Storage 2023, 73, 109133. [Google Scholar] [CrossRef]
  16. Karaca, E.; Gökcen, D.; Pekmez, N.Ö.; Pekmez, K. Electrochemical synthesis of PPy composites with nanostructured MnOx, CoOx, NiOx, and FeOx in acetonitrile for supercapacitor applications. Electrochim. Acta 2019, 305, 502–513. [Google Scholar] [CrossRef]
  17. Baro, M.; Nayak, P.; Baby, T.T.; Ramaprabhu, S. Green approach for the large-scale synthesis of metal/metal oxidenanoparticle decorated multiwalled carbon nanotubes. J. Mater. Chem. A 2013, 1, 482–486. [Google Scholar] [CrossRef]
  18. Dhanabalan, K.; Bhosale, M.; Murugan, N.; Aruchamy, K.; Sriram, G.; Sadhasivam, T.; Oh, T.H. A versatile heterostructure junction of NiO–C-MoOx (X = 2 & 3) composite electrocatalysts for hydrogen evolution reaction. J. Solid. State Chem. 2024, 335, 124702. [Google Scholar] [CrossRef]
  19. Comer, B.M.; Li, J.; Abild-Pedersen, F.; Bajdich, M.; Winther, K.T. Unraveling Electronic Trends in O* and OH* Surface Adsorption in the MO2 Transition-Metal Oxide Series. J. Phys. Chem. C 2022, 126, 7903–7909. [Google Scholar] [CrossRef]
  20. Gryglewicz, G.; Machnikowski, J.; Lorenc-Grabowska, E.; Lota, G.; Frackowiak, E. Effect of pore size distribution of coal-based activated carbons on double layer capacitance. Electrochim. Acta 2005, 50, 1197–1206. [Google Scholar] [CrossRef]
  21. Bardestani, R.; Patience, G.S.; Kaliaguine, S. Experimental methods in chemical engineering: Specific surface area and pore size distribution measurements—BET, BJH, and DFT. Can. J. Chem. Eng. 2019, 97, 2781–2791. [Google Scholar] [CrossRef]
  22. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K.S. Physisorption of gases, with special reference to the evaluation of surface area and pore size distribution (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1051–1069. [Google Scholar] [CrossRef]
  23. Tan, Y.; Xu, X.; Li, Q.; Chen, X.; Che, Q.; Chen, Y.; Long, Y. Constructing ultrathin FeS/FeO H@Fe nano-sheets for highly efficient oxygen evolution reaction. J. Colloid. Interface Sci. 2021, 594, 575–583. [Google Scholar] [CrossRef] [PubMed]
  24. Sudarsanam, P.; Hillary, B.; Amin, M.H.; Hamid, S.B.A.; Bhargava, S.K. Structure-activity relationships of nanoscale MnOx/CeO2 heterostructured catalysts for selective oxidation of amines under eco-friendly conditions. Appl. Catal. B 2016, 185, 213–224. [Google Scholar] [CrossRef]
  25. Kandel, M.R.; Pan, U.N.; Dhakal, P.P.; Ghising, R.B.; Nguyen, T.T.; Zhao, J.; Kim, N.H.; Lee, J.H. Unique heterointerface engineering of Ni2P–MnP nanosheets coupled Co2P nanoflowers as hierarchical dual-functional electrocatalyst for highly proficient overall water-splitting. Appl. Catal. B 2023, 331, 122680. [Google Scholar] [CrossRef]
  26. Murphy, E.; Sun, B.; Rüscher, M.; Liu, Y.; Zang, W.; Guo, S.; Chen, Y.H.; Hejral, U.; Huang, Y.; Ly, A.; et al. Synergizing Fe2O3 Nanoparticles on Single Atom Fe-N-C for Nitrate Reduction to Ammonia at Industrial Current Densities. Adv. Mater. 2024, 36, 2401133. [Google Scholar] [CrossRef]
  27. Quílez-Bermejo, J.; Daouli, A.; Dalí, S.G.; Cui, Y.; Zitolo, A.; Castro-Gutiérrez, J.; Emo, M.; Izquierdo, M.T.; Mustain, W.; Badawi, M.; et al. Electron Transfer from Encapsulated Fe3C to the Outermost N-Doped Carbon Layer for Superior ORR. Adv. Funct. Mater. 2024; early view. [Google Scholar] [CrossRef]
  28. He, Y.; Zhang, J.; Zhou, H.; Yao, G.; Lai, B. Synergistic multiple active species for the degradation of sulfamethoxazole by peroxymonosulfate in the presence of CuO@FeOx@Fe0. Chem. Eng. J. 2020, 380, 122568. [Google Scholar] [CrossRef]
  29. Hu, G.-L.; Hu, R.; Liu, Z.-H.; Wang, K.; Yan, X.-Y.; Wang, H.-Y. Tri-functional molecular relay to fabricate size-controlled CoOx nanoparticles and WO3 photoanode for an efficient photoelectrochemical water oxidation. Catal. Sci. Technol. 2020, 10, 5677–5687. [Google Scholar] [CrossRef]
  30. Shi, H.; Sun, X.Y.; Liu, Y.; Zeng, S.P.; Zhang, Q.H.; Gu, L.; Wang, T.H.; Han, G.F.; Wen, Z.; Fang, Q.R.; et al. Multicomponent Intermetallic Nanoparticles on Hierarchical Metal Network as Versatile Electrocatalysts for Highly Efficient Water Splitting. Adv. Funct. Mater. 2023, 33, 2214412. [Google Scholar] [CrossRef]
  31. Shi, H.; Dai, T.Y.; Sun, X.Y.; Zhou, Z.L.; Zeng, S.P.; Wang, T.H.; Han, G.F.; Wen, Z.; Fang, Q.R.; Lang, X.Y.; et al. Dual-Intermetallic Heterostructure on Hierarchical Nanoporous Metal for Highly Efficient Alkaline Hydrogen Electrocatalysis. Adv. Mater. 2024; early view. [Google Scholar] [CrossRef]
  32. Ehsan, M.A.; Ali, A.; Khan, M.S.; Younas, M.; Zubair, M.; Habib, A.; Hakeem, A.S.; Iqbal, N. Effectual Water Oxidation Reinforced by Three-Dimensional (3D) MnO: A Highly Sustainable Electrocatalyst. ACS Appl. Energy Mater. 2023, 6, 7156–7168. [Google Scholar] [CrossRef]
  33. Jing, T.; Zhang, N.; Zhang, C.; Mourdikoudis, S.; Sofer, Z.; Li, W.; Li, P.; Li, T.; Zuo, Y.; Rao, D. Improving C-N-FeOx Oxygen Evolution Electrocatalysts through Hydroxyl-Modulated Local Coordination Environment. ACS Catal. 2022, 12, 7443–7452. [Google Scholar] [CrossRef]
  34. Han, L.; Dong, S.; Wang, E. Transition-Metal (Co, Ni, and Fe)-Based Electrocatalysts for the Water Oxidation Reaction. Adv. Mater. 2016, 28, 9266–9291. [Google Scholar] [CrossRef] [PubMed]
  35. Babar, P.T.; Lokhande, A.C.; Gang, M.G.; Pawar, B.S.; Pawar, S.M.; Kim, J.H. Thermally oxidized porous NiO as an efficient oxygen evolution reaction (OER) electrocatalyst for electrochemical water splitting application. J. Ind. Eng. Chem. 2018, 60, 493–497. [Google Scholar] [CrossRef]
Figure 1. (a) XRD diffraction pattern, (b) Raman spectra, (c) BET surface area, and (d) pore diameter distribution from BET analysis of as-synthesized MnOx-MoO3, FeOx-MoO3, CoOx-MoO3, and NiOx-MoO3.
Figure 1. (a) XRD diffraction pattern, (b) Raman spectra, (c) BET surface area, and (d) pore diameter distribution from BET analysis of as-synthesized MnOx-MoO3, FeOx-MoO3, CoOx-MoO3, and NiOx-MoO3.
Inorganics 12 00241 g001
Figure 2. XPS spectra of synthesized materials: (a) Mn 2p, (b) Fe 2p, (c) Co 2p, and (d) Ni 2p.
Figure 2. XPS spectra of synthesized materials: (a) Mn 2p, (b) Fe 2p, (c) Co 2p, and (d) Ni 2p.
Inorganics 12 00241 g002
Figure 3. HRTEM and elemental mapping images: (a,b) image of MnOx-MoO3; (cf) C, O, Mn, and Mo mapping of MnOx-MoO3 (scale bar is 400 nm); (g,h) image of FeOx-MoO3; and (il) C, O, Fe, and Mo mapping of FeOx-MoO3.
Figure 3. HRTEM and elemental mapping images: (a,b) image of MnOx-MoO3; (cf) C, O, Mn, and Mo mapping of MnOx-MoO3 (scale bar is 400 nm); (g,h) image of FeOx-MoO3; and (il) C, O, Fe, and Mo mapping of FeOx-MoO3.
Inorganics 12 00241 g003
Figure 4. HRTEM and elemental mapping images: (a,b) image of CoOx-MoO3; (cf) C, O, Co, and Mo mapping of CoOx-MoO3 (scale bar is 400 nm); (g,h) image of NiOx-MoO3; and (il) C, O, Fe, and Ni mapping of NiOx-MoO3.
Figure 4. HRTEM and elemental mapping images: (a,b) image of CoOx-MoO3; (cf) C, O, Co, and Mo mapping of CoOx-MoO3 (scale bar is 400 nm); (g,h) image of NiOx-MoO3; and (il) C, O, Fe, and Ni mapping of NiOx-MoO3.
Inorganics 12 00241 g004
Figure 5. (a) LSV curves recorded in 1.0 M KOH at scan rate 5 mV s−1; (b) electrochemical impedance spectra of all active catalysts; (c) Tafel plot from OER LSV curve; (d) comparative overpotential for OER.
Figure 5. (a) LSV curves recorded in 1.0 M KOH at scan rate 5 mV s−1; (b) electrochemical impedance spectra of all active catalysts; (c) Tafel plot from OER LSV curve; (d) comparative overpotential for OER.
Inorganics 12 00241 g005
Figure 6. (a) Electric double-layer capacitance and (b) electrochemical surface area from Cdl data.
Figure 6. (a) Electric double-layer capacitance and (b) electrochemical surface area from Cdl data.
Inorganics 12 00241 g006
Figure 7. (a) Chrono-potentiometric curves of the CoOx-MoO3 and (b) XRD pattern of carbon cloth and before and after stability test of CoOx-MoO3/CC.
Figure 7. (a) Chrono-potentiometric curves of the CoOx-MoO3 and (b) XRD pattern of carbon cloth and before and after stability test of CoOx-MoO3/CC.
Inorganics 12 00241 g007
Table 1. Comparison of the electrocatalytic activity of metal oxide composites with MoO3 as OER catalysts.
Table 1. Comparison of the electrocatalytic activity of metal oxide composites with MoO3 as OER catalysts.
CatalystPreparation MethodElectrolyte (M)Over Potential @ 10 mA cm−2Reference
MnOThermal annealing1.0 KOH394 mV[32]
FeOxHydrothermal 1.0 KOH379 mV[33]
Co3O4Thermal annealing1.0 KOH490 mV[34]
NiOThermal annealing1.0 KOH330 mV[35]
MnOx-MoO3Thermal annealing1.0 KOH390 mVThis work
FeOx-MoO3Thermal annealing1.0 KOH350 mVThis work
CoOx-MoO3Thermal annealing1.0 KOH310 mVThis work
NiOx-MoO3Thermal annealing1.0 KOH340 mVThis work
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

Dhanabalan, K.; Bhosale, M.; Sriram, G.; Sadhasivam, T.; Oh, T.H. An Investigation of the Interface between Transition Metal Oxides (MnOx, FeOx, CoOx and NiOx)/MoO3 Composite Electrocatalysts for Oxygen Evolution Reactions. Inorganics 2024, 12, 241. https://doi.org/10.3390/inorganics12090241

AMA Style

Dhanabalan K, Bhosale M, Sriram G, Sadhasivam T, Oh TH. An Investigation of the Interface between Transition Metal Oxides (MnOx, FeOx, CoOx and NiOx)/MoO3 Composite Electrocatalysts for Oxygen Evolution Reactions. Inorganics. 2024; 12(9):241. https://doi.org/10.3390/inorganics12090241

Chicago/Turabian Style

Dhanabalan, Karmegam, Mrunal Bhosale, Ganesan Sriram, Thangarasu Sadhasivam, and Tae Hwan Oh. 2024. "An Investigation of the Interface between Transition Metal Oxides (MnOx, FeOx, CoOx and NiOx)/MoO3 Composite Electrocatalysts for Oxygen Evolution Reactions" Inorganics 12, no. 9: 241. https://doi.org/10.3390/inorganics12090241

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