Next Article in Journal
Mechanism and Kinetic Study on Synthesis of Methacrolein Catalyzed by Amine/Acid
Next Article in Special Issue
Fabricating Dispersed Fine Silver Nanoparticles on Liquid Substrate for Improved Photocatalytic Water Splitting Efficiency
Previous Article in Journal
Effect of Ce Content on the Chemical Looping Oxidative Dehydrogenation of Propane to Propylene over a VOx-CeO2/γ-Al2O3 Oxygen Carrier
Previous Article in Special Issue
PGM-Free Electrocatalytic Layer Characterization by Electrochemical Impedance Spectroscopy of an Anion Exchange Membrane Water Electrolyzer with Nafion Ionomer as the Bonding Agent
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tuning Electrochemical Hydrogen-Evolution Activity of CoMoO4 through Zn Incorporation

by
Sanaz Chamani
1,†,
Ebrahim Sadeghi
1,2,†,
Ugur Unal
3,4,
Naeimeh Sadat Peighambardoust
1 and
Umut Aydemir
1,4,*
1
Koç University Boron and Advanced Materials Applications and Research Center (KUBAM), Sariyer, Istanbul 34450, Turkey
2
Graduate School of Sciences and Engineering, Koç University, Sariyer, Istanbul 34450, Turkey
3
Koç University Surface Science and Technology Center (KUYTAM), Sariyer, Istanbul 34450, Turkey
4
Department of Chemistry, Koç University, Sariyer, Istanbul 34450, Turkey
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2023, 13(5), 798; https://doi.org/10.3390/catal13050798
Submission received: 10 April 2023 / Revised: 21 April 2023 / Accepted: 21 April 2023 / Published: 24 April 2023
(This article belongs to the Special Issue Electrocatalytic Hydrogen Evolution Reaction through Water Splitting)

Abstract

:
Designing cheap, efficient, and durable electrocatalysts on three-dimensional (3D) substrates such as nickel foam (NF) for the hydrogen-evolution reaction (HER) is in high demand for the practical application of electrochemical water splitting. In this work, we adopted a simple one-step hydrothermal method to realize the incorporation of Zn into the lattice of CoMoO4 with various atomic concentrations—Co1-xZnxMoO4 (x = 0, 0.1, 0.3, 0.5, and 0.7). The morphological studies demonstrated that parent CoMoO4 consists of nanoflowers and nanorods. However, as the concentration of Zn increases within the host CoMoO4, the portion of nanoflowers decreases and simultaneously the portion of nanorods increases. Moreover, the substitution of Zn2+ in place of Co2+/Co3+ creates oxygen vacancies in the host structure, especially in the case of Co0.5Zn0.5MoO4, giving rise to lower charge-transfer resistance and a higher electrochemically active surface area. Therefore, among the prepared samples, Co0.5Zn0.5MoO4 on NF showed an improved HER performance, reaching 10 mA cm−2 at an overpotential as low as 204 mV in a 1.0 M KOH medium. Finally, the Co0.5Zn0.5MoO4 electrode exhibited robust long-term stability at an applied current density of 10 mA cm−2 for 20 h. The Faradaic efficiency determined by a gas chromatograph found that the hydrogen-production efficiency varied from 94% to 84%.

Graphical Abstract

1. Introduction

The universally rapid consumption of traditional fossil fuels and the pernicious environmental impacts they create have encouraged researchers to actively and urgently seek a sustainable, renewable, and environmentally friendly energy source as an alternative [1,2]. Electrocatalytic water splitting is a potential approach to convert renewable electricity—generated from renewable energy resources—to hydrogen as a pure, viable, and carbon dioxide-free energy source [3,4]. The efficiency of the water-splitting process is closely dependent on the degree of competence of the electrocatalysts to drive the evolution of the hydrogen (HER) and oxygen (OER) from the medium [5,6]. Noble metal-based electrocatalysts, despite their robust activity, are not sufficiently durable. Additionally, these materials are expensive and rare [7]. Therefore, to circumvent these issues, great efforts have been made to design and develop catalysts based on earth-abundant transition metals (TMs) (e.g., Ni, Co, Fe, Zn, Cu, Mo) [8]. A lot of recent publications have proved the excellent performance of TM-based compounds in water electrolysis. Among these compounds, transition metal oxides (TMOs), because of their low cost, rich redox properties, abundance, and durability under alkaline conditions, are one of the most promising candidates to consider [9,10].
Recently, there has been a surge of interest in the electrochemical properties of bimetallic oxides in comparison to single-component oxides due to their improved characteristics [11,12,13]. Bimetallic oxide catalysts could alter the electronic structures of the catalysts by regulating the composition, physical, and chemical aspects of the metals in the structure [14,15]. Moreover, bimetallic oxides offer multiple active sites, trap inactive active sites, and elevate the conductivity at the same time [14,16]. TM molybdates, as a subgroup of bimetallic compounds, have received much attention because of their high electrochemical activities, which can be ascribed to their rich polymorphism and high conductivity originating from metal molybdates [12,17]. CoMoO4 is considered a potentially good electrocatalytic material, owing to the synergy between the redox ability of Co and the high adsorption capacity of Mo in its framework [12,18,19,20]. Yet, CoMoO4 shows limited HER performance in alkaline media, which is most probably attributed to its poor intrinsic conductivity. Thus, it is a significant task to modify the electrocatalytic features of CoMoO4 to act effectively in the alkaline HER [21].
Multiple strategies have been introduced to promote the conductivity and electrocatalytic properties of materials, including phase transition [22,23], coupling with highly conductive carbon materials, doping with heteroatoms [24], strain engineering [21,25], facet control [26], and oxygen vacancies [27,28]. Among them, elemental doping is a common way to optimize the physicochemical properties of catalysts and thereby adjust the electronic structure, which could availably improve HER performance [29,30,31]. Only a few reports have been released that elaborate on the effect of elemental doping on the HER activity of CoMoO4, and those which have done so mostly focused on anion doping. Wen et al. [14] utilized a phosphorous doping technique to fabricate a controlled catalyst P-CoMoO4@Ni with a hierarchical structure with different morphologies on the surface of nickel foam (NF). The best HER performances reported by this group were obtaining current densities of 100 and 150 mA cm−2 under low overpotentials of 98 and 162 mV, respectively. Jiao et al. [32] took advantage of phosphorous doping using a facile hydrothermal method followed by low-temperature phosphidation to modulate the electronic structure of CoMoO4 on NF for an alkaline HER. The developed P-CoMoO4/NF demonstrated electrocatalytic activity toward the HER in 1.0 M KOH with a low overpotential of 89 mV at 10 mA cm−2.
In this work, we report the effect of Zn substitution in the Co sites of monoclinic CoMoO4 on the HER in a 1.0 M KOH medium. Ling et al. [24] conducted a thorough theoretical and experimental investigation on the electronic structure modulation of CoO nanorods by dual Ni and Zn doping for the HER. Based on combined density functional theory (DFT) calculations and careful microscopic and spectroscopic characterizations, they showed that the Zn dopants distribute inside the host oxide and control bulk electronic structure to promote electrical conduction. In addition, introducing a Zn dopant with a slightly larger atomic radius compared to Co into the host, CoMoO4, and different electronic configurations will tailor the structural properties along with the chemical environment. Furthermore, comparing the nearly identical ionic radii of Co2+ (0.063 nm) and Zn2+ (0.074 nm) as well as their similar electronegativities—1.88 for Co and 1.65 for Zn—indicates the suitability of the Zn dopant in the host CoMoO4 [33]. The Zn dopant with varying atomic concentrations was substituted using a simple one-pot hydrothermal technique in the final electrode, Co1-xZnxMoO4, where x = 0.1, 0.3, 0.5, and 0.7. The effect of varied Zn substitution for the alkaline HER is discussed. Among these catalysts, Co0.5Zn0.5MoO4 displayed the highest HER performance, with an overpotential of 204 mV at 10 mA cm−2 and a Tafel slope of 162 mV dec−1 in 1.0 M KOH.

2. Results and Discussion

2.1. Chemical Characterization

The schematic growth of Co1-xZnxMoO4 (x = 0.0, 0.1, 0.3, 0.5, 0.7) on NF can be visualized in Figure 1a (further details can be found in Materials and Methods, Section 3.2). The XRD was performed on the as-prepared materials. The XRD patterns of Co1-xZnx MoO4 and the theoretical pattern of CoMoO4 are illustrated in Figure 1b. The characteristic XRD peaks at 2 θ of 13.10, 18.95, 23.12, 26.40, 28.34, 31.96, 33.47, 36.62, 38.66, 40.10, 43.45, 45.1, and 47.07 correspond to the respective crystalline planes of the (001), ( 2 ¯ 01), (021), (220), ( 3 ¯ 11), (112), ( 2 ¯ 22), (400), (040), (330), ( 4 ¯ 22), ( 5 ¯ 11), and (421) and are well-matched with the monoclinic structure of β -CoMoO4 ((ICDD No. 04-017-6377). It is pertinent to mention that the dashed line around 2 θ of 26.40 ° indicates a noticeable peak shift toward smaller 2 θ values with respect to the increase in the concentration of Zn in the host CoMoO4, confirming the successful substitution of Zn and thereby implying an enlargement in lattice parameters (a better view of the XRD shift for (220) can be found in Figure S1, in the Supplementary Materials). Figure 1c depicts the crystal structure design of β -CoMoO4, which crystallizes in the space group C2/m (No.12). In this scheme, CoO6 octahedrons are in purple and the tetrahedrons MoO4 are in a gray color [34].
An FE-SEM equipped with EDS, TEM, and HR-TEM was employed to study the surface morphology, analyze the elemental composition, and identify the interplanar spacing of as-prepared materials. Figure 2a–e displays the corresponding SEM images of pure CoMoO4 and Zn-substituted CoMoO4. The morphology of the catalysts grown on the NF substrate is a combination of two structures of nanoflowers and nanorods. The SEM microstructure of CoMoO4 in Figure 2a demonstrates that the portion of nanoflowers and nanorods on the NF are somehow equal. However, the comparison of the SEM image of Co0.9Zn0.1MoO4 in Figure 2b with that of CoMoO4 indicates that the size of nanoflowers reduced. Moreover, as the concentration of Zn increases within the host CoMoO4, the portion of nanoflowers decreases and simultaneously the portion of nanorods increases. With this in mind, materials containing a higher concentration of Zn mostly consist of nanorods (refer to high-magnification SEM photographs shown in Figure 2d,e). The diameter of the nanorods was estimated to be between 100 and 300 nm. However, the length of the nanorods can reach a couple of hundreds of micrometers. The uniform growth of all as-obtained catalysts on the NF can be seen in Figure S2. Additionally, the main ingredients of the prepared materials (Mo, Co, O, and Zn) were verified through EDS analysis (the details can be found in Figure S3 and Table S1).
The observation of the TEM micrographs indicated that for the Co0.5Zn0.5MoO4 sample, the dominating morphology is the nanorod-like structure (Figure 2f). The HR-TEM image of Co0.5Zn0.5MoO4 revealed that the interplanar spacing of the Co0.5Zn0.5MoO4 nanorods is about 0.46 nm (Figure 2f, fourth image from left) and 0.384 nm (Figure 2f, fifth image from left), which correspond to the ( 2 ¯ 01) and (021) planes of monoclinic CoMoO4, respectively. Furthermore, the HAADF-STEM image taken from the selected sample confirmed the even distribution of Co, Zn, Mo, and O on the random nanorods (see Figure 2g).
Fourier transform infrared (FT-IR) characterization was utilized to further verify the successful incorporation of Zn into Co sites in CoMoO4. Figure S4 displays the FT-IR spectra of all samples. The absorption signals that emerged at 928 and 783 cm−1 denote the stretching vibration mode of the O–Mo–O bond [35,36,37,38]. The peak located at 696 cm−1 was attributed to the vibration of the Co–O–Mo group [37]. The band that appeared at 408 cm−1 represents the Mo–O and CoO6 groups of CoMoO4 [39,40]. The absorption bands in the region of 500–1000 cm−1 demonstrate shifts in the wavenumber as the content of Zn changes in the sample. This shift toward higher wavenumber values—blueshift—is an indication of the substitution of Zn2+ for Co2+, which increases monotonically with the increase of Zn concentration [41,42]. As a result, the incorporation of Zn atoms into the CoMoO4 lattice strengthens the metal–oxygen bonds in the structure.
The surface chemical composition and oxidation states of prepared catalysts were characterized by XPS. The presence of expected elements was verified by the XPS survey shown in Figure S5a. The high-resolution XPS spectra of CoMoO4 and Co0.5Zn0.5MoO4 are depicted in Figure 3. From Figure 3a, we note how the Co 2p spectrum in CoMoO4 revealed two major doublets and their corresponding shake-up satellite peaks. The lower BEs centered at 780.68 and 784.59 eV can be related to Co 2p3/2, while those located at higher BEs (796.71 and 800.20 eV) are correlated with Co 2p1/2. The Co 2p peak analysis shows that Co exists in oxidation states of 2+ and 3+ [9]. The high-resolution Mo 3d spectrum can be divided into two sets of doublets, which are ascribed to Mo 3d3/2 at higher BEs (235.03 and 236.79 eV) and Mo 3d5/2 at lower BEs (231.81 and 233.77 eV). These peaks confirm the presence of both oxidation states of Mo6+ and Mo4+ [43] on the surface. The O 1s XPS spectrum for CoMoO4 also displayed three peaks at BEs—530.14, 533.54, and 537.06 eV—which can be attributed to the metal–oxygen, oxygen vacancy, and adsorbed hydroxyl groups on the surface [44,45].
The deconvolution of relevant XPS spectra for Co0.5Zn0.5MoO4 (shown in Figure 3b) exhibited that Zn 2p species are composed of two spin-orbit doublets of Zn 2p3/2 and Zn 2p1/2, suggesting the substitution of Zn2+ into the host structure (refer to Table S2). In addition, an evident peak shift was observed for both Co 2p and Mo 3d toward higher BEs compared with CoMoO4. The substitution of Zn2+ in place of Co2+/Co3+ gives rise to a mitigation of positive charge, resulting in the formation of more oxygen vacancies to maintain the charge neutrality in the structure of Co0.5Zn0.5MoO4 and therefore higher oxidation states on the metal sites [46]. It can be concluded that the Zn-site stabilizes the Co-site by oxidizing the Co to receive an improved structural sturdiness without experiencing significant deformation [46,47]. Similar to CoMoO4, the core-level O 1s XPS spectrum for Co0.5Zn0.5MoO4 was analyzed and three peaks were obtained at 530.07, 532.17, and 534.12 eV, corresponding to the identical oxygen-bearing groups. As expected, the amount of oxygen vacancy increased in Co0.5Zn0.5MoO4 relative to unsubstituted CoMoO4. The XPS spectra of all elements for each sample can be observed in Figure S5b–e. The subtle dissimilarity between the O 1s spectra of Co0.5Zn0.5MoO4 and other electrodes stems from the larger oxygen vacancy peak for Co0.5Zn0.5MoO4. It is worth noting that a correlation exists between the concentration of Zn substitution and the degree of XPS shift toward higher BEs for the Co 2p, Mo 3d, and Zn 2p spectra in all samples except for Co0.3Zn0.7MoO4. As the concentration of Zn increases from x = 0.1 to x = 0.5, the XPS peaks shift to higher BEs compared to CoMoO4 without substitution. Among the samples, Co0.9Zn0.1MoO4 and Co0.5Zn0.5MoO4 exhibited the least and most peak shift in the direction of higher BEs, respectively. Nonetheless, the trend is disrupted by the increase of Zn concentration to x = 0.7. It can be calculated that until x = 0.5, the structure possesses the highest amount of oxygen vacancy with high structural integrity. The creation of oxygen vacancies can modulate the electronic structure, boost conductivity, and enhance catalytic performance [48].

2.2. Electrocatalytic Hydrogen Evolution

The electrochemical properties of as-prepared catalysts for the HER were examined on NF with a ~1 mg cm−2 catalyst loading as a working electrode in 1.0 M KOH solution.
Moreover, Figure 4a displays the electrocatalytic HER activities of as-prepared materials on the NF substrate along with blank NF and commercial Pt/C (20 wt%) as a comparison. Based on the results, the overpotentials (measured at 10 mA cm−2) of the materials exhibit the following trend: Pt/C < Co0.5Zn0.5MoO4 < Co0.3Zn0.7MoO4 < Co0.7Zn0.3MoO4 < Co0.9Zn0.1MoO4 < CoMoO4 < blank NF (Figure 4b). A quick examination of this trend shows that the state-of-the-art Pt/C@NF electrode and blank NF, respectively, featured the lowest and highest overpotential at 10 mA cm−2. Among the as-prepared materials, CoMoO4 has the lowest performance and the highest overpotential. On the other hand, Co0.5Zn0.5MoO4 showcased the best performance and the lowest overpotential. From the results it can be concluded that Zn incorporation modified the HER electrocatalysis, explaining that all Zn-substituted materials have better activity compared to their unsubstituted counterpart. To compare the results obtained in this work, the HER performances of previous studies have been tabulated in Table S3 [18,20,49,50,51,52,53,54,55].
To explore the electrochemical kinetic behavior, the Tafel slopes were calculated. The Tafel slopes of as-obtained materials are depicted in Figure 4c. The HER kinetics, however, are facile irrespective of the specified overpotential of the catalyst. In this sense, it can be inferred that all reported catalysts have quite identical HER kinetics [9]. In addition, the kinetics of the HER in alkaline media can be realized by the mechanisms of adsorbing and desorbing hydrogen atoms or molecular hydrogen via either the Volmer–Heyrovsky or Volmer–Tafel pathways [56]. The underlying reactions are:
H2O + e + * → Hads* + OH (Volmer step)
Hads* + H2O + e → H2 + OH + * (Heyrovsky step)
Hads* + Hads* → H2 + 2* (Tafel step)
where * denotes the available active sites and Hads* indicates the presence of atomic hydrogen at the active site. To understand the mechanism of the HER, Tafel plots obtained from LSV curves can be used as a means of inference.
The reported Tafel slopes for the Volmer, Heyrovsky, and Tafel steps are 120, 40, and 30 mV dec−1, respectively. Among the prepared samples, the commercial Pt/C catalyst shows the smallest Tafel slope value (113 mV dec−1), suggesting that the Volmer reaction governs the overall HER process [57]. According to Figure 4c, all as-prepared materials showed Tafel slope values beyond 120 mV dec−1, implying that the rate of the reaction is determined by the adsorption of hydrogen atoms on the active sites. The Co0.5Zn0.5MoO4 sample showed a lower Tafel slope (162.7 mV dec−1), signifying faster kinetics for hydrogen adsorption at active sites compared to other samples. In order to examine the reason for the high activity of Co0.5Zn0.5MoO4, the ECSA was estimated with respect to Cdl for these materials [55]. As there is a direct relation between ECSA and Cdl, the higher the Cdl value, the better the electrocatalytic activity. The Co0.5Zn0.5MoO4 gave the maximum Cdl value (5.35 mF cm−2), which is consistent with the HER activity (refer to Figure 4d). Therefore, the larger electroactive surface area is a favorable parameter for enhancing the electron transfer rate. The recorded CV scans within the potential window of 0.75–0.85 V vs. RHE for the as-obtained catalysts can be visualized in Figure S6.
To supplement the data from the electroactive surface area, we analyzed the BET surface area. The results showed that even though the BET values are relatively comparable, Co0.5Zn0.5MoO4 possesses the highest BET surface area (10.95 m2/g). The higher surface area offers more available active sites for the adsorption of hydrogen atoms, encourages fast charge transfer, and in turn, boosts electrocatalysis [58]. The typical N2 adsorption–desorption isotherms are presented in Figure S7. Further details derived from BET analysis are tabulated in Table S4. To further unmask the HER kinetics, EIS measurements were implemented. Figure 4e presents the Nyquist plots of all fabricated samples on the NF substrate. The smallest and largest semicircles belong to Co0.5Zn0.5MoO4 and CoMoO4, which describe the degree of charge-transfer resistance (RCt) between electrolyte and electrode. This means that charge transfer favored the Co0.5Zn0.5MoO4, leading to a facilitated kinetic barrier, and in turn improved electrocatalytic performance [59]. It must be pointed out that the inset image of Figure 4e symbolizes the equivalent circuit model for the charge-transfer path on the catalysts. In this model, Rs denotes resistance induced by the solution and Rp indicates resistance coming from the porous substrate. In addition, to fulfill the non-ideal behavior of the capacitive elements, a constant phase element (CPE) was applied. Similarly, the double-layer capacitance in the interface of electrode/solution was taken into account through Cdl. Further details associated with the charge transfer on the catalysts can be found in Table S5.
From a perspective of practical application, the HER long-term durability of Co0.5Zn0.5MoO4 was evaluated by performing chronopotentiometry measurement at the constant current density of 10 mA cm−2 in 1.0 M KOH. As presented in Figure 4f, no observable degradation was discerned after 20 h of continuous HER, demonstrating the robust tolerance of Co0.5Zn0.5MoO4 in an alkaline solution. It is worth noting that the potential decreases within the first 2 h, which might be on account of changes in the oxidation states of transition metals and the exfoliation of nanostructures on the surface, giving rise to abundant active sites on the electrode. The LSV curve after 20 h disclosed that the HER overpotential at 10 mA cm−2 increased from 204 mV to 248 mV (Figure 4g). Moreover, the long-term stability of Co0.5Zn0.5MoO4 was assessed by 1000 CV scans over the reduction potential window. The HER after 1000 CV cycles was obtained as 236 mV at 10 mA cm−2, confirming its favorable endurance (Figure 4h).

2.3. Post Electrolysis

To spot any changes in the morphology and chemical composition on the surface after long-term stability, the Co0.5Zn0.5MoO4 electrode was checked by FE-SEM, TEM, and XPS. Figure 5a shows the post-electrocatalysis FESEM image of the catalyst on the NF. From the high-magnification SEM photograph, we can see that the nanorod morphology of the catalyst remained largely unchanged. However, the smooth nanorods transformed into a rough-shaped structure. Moreover, Figure 5b demonstrates the TEM and HR-TEM images. The post-electrolysis electrode mostly consists of nanosheets with an interplanar spacing of 0.46 nm associated with CoMoO4 ( 2 ¯ 01), confirming there is no phase transformation in the catalyst after 20 h of HER. The same observation was substantiated by the SAED image shown in the inset of Figure 5b. Finally, Figure 5c presents the high-resolution Co 2p, Mo 3d, O1s, and Zn 2p XPS spectra after HER stability characterization. All peaks that appeared for all elements shifted to lower values of BE compared with the fresh sample (refer to Table S2 and Figure 3b). The comparison of the atomic proportion of elements on the surface before and after the durability test indicated that the amount of Co 2p (6.03% → 4.98%), Mo 3d (10.28% → 0.51%), and Zn 2p (4.3% → 0.72%) decreased meaningfully. On the contrary, the atomic proportion of O 1s after the HER stability measurement elevated to 65.01% compared to the fresh sample (51.03%). From these observations, it can be concluded that the oxidation states on the metals reduced after 20 h of continuous HER. Moreover, Mo and Zn in the vicinity of multiple oxygen defects might be more favorable metal sites for the adsorption of OH compared to Co. The schematic illustration of the HER on the surface of the Co0.5Zn0.5MoO4 electrode can be observed in Figure 6a.
The amount of H2 production in a three-electrode configuration was measured at a fixed current density of 10 mA cm−2 for 150 min at a regular interval of 30 min by a gas chromatograph, as shown in Figure 6b. The Faradaic efficiency (FE) for the HER was estimated via FE = n e x p Q / Z F , in which n e x p denotes the number of moles of H2 experimentally produced at applied current density, Q (C) indicates the total amount of charge, Z number of transferred electrons (i.e., in HER, Z = 2 ), and F is the Faraday’s constant (96,485 C mol−1). The hydrogen-production efficiency started from 94% in the first measurement and dropped to 84% in the fifth measurement (Figure 6c). The loss in Faradaic efficiency can be attributed to the partial detachment of the catalyst from the NF substrate after 150 min under exposure to 10 mA cm−2.

3. Materials and Methods

3.1. Materials

Sodium molybdate dihydrate [Na2MoO4·2H2O, Sigma-Aldrich, ACS reagent, ≥99%], cobalt nitrate hexahydrate [Co(NO3)2·6H2O, Sigma-Aldrich, 99.999% trace metal basis], and zinc nitrate hexahydrate [Zn(NO3)2·6H2O, Sigma-Aldrich, 98%] were purchased and used without any prior treatment.

3.2. Preparation of Co1-xZnxMoO4@NF

A piece of commercial NF (1.5 × 3 cm−2) was cleaned ultrasonically with 2.0 M HCl for 10 min to remove metal oxide layers on the surface and then rinsed with ethanol, acetone, and deionized (DI) water, consecutively, each for 10 min. The growth of Co1-xZnxMoO4 (x = 0.0, 0.1, 0.3, 0.5, 0.7) on NF was performed using a facile hydrothermal method (refer to Figure 1a). For the preparation of bare CoMoO4 (x = 0), 2 mmol NaMoO4·2H2O and 2 mmol Co(NO3)2·6H2O were completely dissolved in 40 mL DI water through stirring to receive a red-colored transparent solution. The obtained solution was transferred into a Teflon-lined autoclave with a piece of NF and heated at 140 °C for 6 h. Once the reactor cooled down naturally, the resulting product was collected, washed several times with DI water and ethanol, and vacuum-dried at 60 °C for 24 h. Finally, the purple powders were calcined at 350 °C for 2 h to acquire β -CoMoO4. For the substitution of Co with Zn, a similar procedure was applied. In this regard, varying concentrations of Zn(NO3)2·6H2O were introduced to adjust the molar ratio between Co and Zn. Table S6 summarizes the defined molar ratios of initial materials for the fabrication of Co1-xZnxMoO4.

3.3. Preparation of Pt/C@NF Electrocatalyst

According to our previous study [9], 1 mg commercial 20% Pt/C was ultrasonically dispersed in a solution of pure ethanol (300 μ L ) and DI water (200 μ L ) for 30 min. Afterward, 10 μ L Nafion solution was added and sonicated for another 30 min to gain a homogeneous ink solution. The prepared ink was dipped at once on a piece of NF to obtain a mass loading of approximately 1 mg cm−2.

3.4. Materials Characterization

The crystal phase and purity of the products were analyzed by X-ray diffraction (XRD, Rigaku Mini Flex 600, Cu Kα radiation, λ = 1.5418 Å, Rigaku, Tokyo, Japan). The morphology was investigated by a field emission-scanning electron microscope (FE-SEM; Zeiss Ultra Plus, Zeiss, Jena, Germany) equipped with an energy-dispersive X-ray spectroscopy detector (EDS, Bruker Xflash 5010, 123 eV spectral resolution, Bruker Corporation, Billerica, MA, USA). The microstructure, selected area electron diffraction (SAED), and lattice fringes were further studied by high-resolution transmission electron microscopy (HR-TEM; Thermo Scientific Talos F200S 200 kV, Thermo Fisher Scientific, Waltham, MA, USA). In addition, high-angle annular dark field-scanning transmission electron microscopy (HAADF-STEM) images and the associated EDS-STEM mappings were taken on a microscope (Hitachi HF5000 200kV (S)TEM, working at HR mode, Brisbane, Australia). The surface composition and oxidation states were obtained by X-ray photoelectron spectroscopy (XPS, Thermo Scientific K-Alpha with an Al Kα monochromator source, 1486.6 eV, Thermo Fisher Scientific, Waltham, MA, USA). All XPS spectra were corrected according to the binding energy (BE) of C 1s = 284.50 eV. The functional groups were identified by Fourier transform-infrared spectroscopy (FT-IR, JASCO 6800 full vacuum and FT-IR microscope, Jasco Corporation, Tokyo, Japan). The BET-specific surface area was determined through N2 adsorption–desorption isotherms (BET, Micromeritics ASAP 2010, Micromeritics Instruments Corporation, Norcross, GA, USA).

3.5. Electrochemical Measurements

The electrochemical properties of the prepared catalysts were investigated by conducting tests on an Autolab potentiostat/galvanostat instrument using a standard three-electrode cell. The cell consisted of a reversible hydrogen electrode (RHE; HydroFlex), Pt spring, and as-prepared catalysts on NF (0.5 cm × 1 cm) as reference, counter, and working electrodes, respectively, under 1.0 M KOH medium. To this end, the linear sweep voltammetry (LSV) curves were recorded at a scan rate of 5 mV s−1 from 0 to −1 V (vs RHE) toward the HER. The required overpotential for the HER was calculated at a current density of 10 mA cm−2. The cycling behavior of the catalysts was examined by recording cyclic voltammetry (CV) for 1000 cycles at a scan rate of 100 mV s−1. The electrochemically active surface area (ECSA) was acquired through the double-layer capacitance (Cdl) with respect to the CV measured at various scan rates of 0.04, 0.06, 0.08, 0.1, 0.12, 0.14, 0.16, 0.18, and 0.2 V s−1 across the potential window from 0.75 to 0.85 V versus RHE. The long-term stability of the best-performing catalyst was evaluated by chronopotentiometry test at 10 mA cm−2 for 20 h. In addition, a frequency ranging from 100 kHz to 0.1 Hz with a 10 mV RMS sinusoidal modulation at 0 V was utilized to obtain the EIS curves. Finally, H2 gas production for the three-electrode cell water electrolysis was detected and quantified using a gas chromatograph (7820A, GC-System, Agilent, Agilent Technologies, Santa Clara, CA, USA) equipped with a thermal conductivity detector (TCD). The experiment was carried out in a well-sealed homemade system to guarantee there was no gas leakage.

4. Conclusions

In brief, Zn was successfully introduced into the lattice of monoclinic CoMoO4 via a one-pot hydrothermal protocol. The incorporation of Zn modulated the electronic structure of the host material, which caused the Co to shift to high oxidation states and stabilized the surface bond by the formation of oxygen vacancies. The electrocatalytic HER performance of Zn-substituted CoMoO4 (Co1-xZnxMoO4 (x = 0.1, 0.3, 0.5, and 0.7)) showed enhanced activity compared to parent CoMoO4. Amidst the as-obtained samples of NF, the electrode with x = 0.5 demonstrated the highest double-layer capacitance (Cdl = 5.35 mF cm−2) and the lowest charge-transfer resistance (Rct = 133 Ω). In addition, Co0.5Zn0.5MoO4 displayed the highest concentration of oxygen vacancy based on XPS results, leading to facilitated electron transfer. Due to the foregoing factors, Co0.5Zn0.5MoO4 featured better hydrogen-evolution activity, affording 10 mA cm−2 at an overpotential of 204 mV. Moreover, both chronopotentiometry and 1000 CV scans confirmed the excellent long-term durability of the catalyst. More importantly, the Faradaic efficiency determined the capacity of hydrogen production on the Co0.5Zn0.5MoO4 electrode to be 10 mA cm−2, which differed from 94% to 84%.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13050798/s1. Figure S1. The shift of XRD (220) peaks for the prepared Co1-xZnxMoO4 samples. Figure S2. Depiction of the homogenous proliferation of catalysts on the backbone of NF substrate, (a) CoMoO4, (b) Co0.9Zn0.1MoO4, (c) Co0.7Zn0.3MoO4, (d) Co0.5Zn0.5MoO4, and (e) Co0.3Zn0.7MoO4 Figure S3. (a–e) SEM/EDS and elemental mappings of CoMoO4, Co0.9Zn0.1MoO4, Co0.7Zn0.3MoO4, and Co0.5Zn0.5MoO4, Co0.3Zn0.7MoO4 on NF, respectively. Table S1. EDS elemental composition of Co1-xZnxMoO4 (x = 0, 0.1, 0.3, 0.5, and 0.7). Figure S4. FT-IR Spectra of prepared Co1-xZnxMoO4. Table S2. The XPS binding energy values for CoMoO4, Co0.5Zn0.5MoO4, and Co0.5Zn0.5MoO4 after HER stability. Figure S5. (a) XPS survey, (b) Co 2p, (c) Mo 3d, (d) O 1s, and (e) Zn 2p for Co1-xZnxMoO4 (x = 0, 0.1, 0.3, 0.5, and 0.7). Table S3. Comparing the HER performance of present work with previous studies. Figure S6. Typical cyclic voltammetry (CV) curves obtained at different scan rates (0.04–0.2 V s−1) within the potential window 0.75–0.85 V vs. RHE for (a) CoMoO4, (b) Co0.9Zn0.1MoO4, (c) Co0.7Zn0.3MoO4, (d) Co0.5Zn0.5MoO4, and (e) Co0.3Zn0.7MoO4. Figure S7. N2 adsorption–desorption isotherms of (a) CoMoO4, (b) Co0.9Zn0.1MoO4, (c) Co0.7Zn0.3MoO4, (d) Co0.5Zn0.5MoO4, and (e) Co0.3Zn0.7MoO4. Table S4. Further details derived from BET analysis of the samples. Table S5. Fit parameters for the samples derived from EIS experiments. Table S6. The defined molar ratios of initial materials for the fabrication of Co1-xZnxMoO4 (x = 0, 0.1, 0.3, 0.5, 0.7).

Author Contributions

S.C. and E.S. synthesized the catalysts and performed the chemical and electrochemical measurements. N.S.P. helped with the interpretation of all analyses. U.U. helped with the Faradaic efficiency experiments and their interpretation. U.A. supervised the whole project from synthesis to interpretation of the results. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the Turkish Academy of Sciences—Outstanding Young Scientist Award Program.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Acknowledgments

The authors would like to acknowledge Barış Yağcı and Amir Motallebzadeh, at Koç University Surface Science and Technology Center (KUYTAM) for their help with characterizations. Likewise, we thank Gülcan Çorapcıoğlu at Koç University Nanofabrication and Nanocharacterization Center for Scientific and Technological Advanced Research (n2STAR) for her help with HAADF-STEM analysis. Finally, we are grateful to Süleyman Tekmen from Bayburt University Central Research Laboratory (BUMER) for the HR-TEM measurements.

Conflicts of Interest

There are no conflicts of interest to declare.

References

  1. Anantharaj, S.; Ede, S.R.; Sakthikumar, K.; Karthick, K.; Mishra, S.; Kundu, S. Recent trends and perspectives in electrochemical water splitting with an emphasis on sulfide, selenide, and phosphide catalysts of Fe, Co, and Ni: A review. ACS Catal. 2016, 6, 8069–8097. [Google Scholar] [CrossRef]
  2. Chamani, S.; Sadeghi, E.; Peighambardoust, N.S.; Doganay, F.; Yanalak, G.; Eroglu, Z.; Aslan, E.; Asghari, E.; Metin, O.; Patir, I.H.; et al. Photocatalytic hydrogen evolution performance of metal ferrites/polypyrrole nanocomposites. Int. J. Hydrogen Energy 2022, 47, 32940–32954. [Google Scholar] [CrossRef]
  3. Feng, C.; Faheem, M.B.; Fu, J.; Xiao, Y.; Li, C.; Li, Y. Fe-based electrocatalysts for oxygen evolution reaction: Progress and perspectives. ACS Catal. 2020, 10, 4019–4047. [Google Scholar] [CrossRef]
  4. Ramírez, A.M.R.; Heidari, S.; Vergara, A.; Aguilera, M.V.; Preuss, P.; Camarada, M.B.; Fischer, A. Rhenium-Based Electrocatalysts for Water Splitting. ACS Mater. Au 2023. [Google Scholar] [CrossRef]
  5. Wang, J.; Yue, X.; Yang, Y.; Sirisomboonchai, S.; Wang, P.; Ma, X.; Abudula, A.; Guan, G. Earth-abundant transition-metal-based bifunctional catalysts for overall electrochemical water splitting: A review. J. Alloys Compd. 2020, 819, 153346. [Google Scholar] [CrossRef]
  6. Kim, J.; Choi, S.; Cho, J.; Kim, S.Y.; Jang, H.W. Toward multicomponent single-atom catalysis for efficient electrochemical energy conversion. ACS Mater. Au 2021, 2, 1–20. [Google Scholar] [CrossRef] [PubMed]
  7. Sun, H.; Xu, X.; Kim, H.; Jung, W.; Zhou, W.; Shao, Z. Electrochemical water splitting: Bridging the gaps between fundamental research and industrial applications. Energy Environ. Mater. 2022, e12441. [Google Scholar] [CrossRef]
  8. Yang, W.-D.; Xiang, J.; Zhao, R.-D.; Loy, S.; Li, M.-T.; Ma, D.-M.; Li, J.; Wu, F.-F. Nanoengineering of ZnCo2O4@CoMoO4 heterogeneous structures for supercapacitor and water splitting applications. Ceram. Int. 2023, 49, 4422–4434. [Google Scholar] [CrossRef]
  9. Sadeghi, E.; Peighambardoust, N.S.; Chamani, S.; Aydemir, U. Designing In Situ Grown Ternary Oxide/2D Ni-BDC MOF Nanocomposites on Nickel Foam as Efficient Electrocatalysts for Electrochemical Water Splitting. ACS Mater. Au 2023, 3, 143–163. [Google Scholar] [CrossRef]
  10. Chamani, S.; Khatamian, M.; Peighambardoust, N.S.; Aydemir, U. Microwave-Assisted Auto-Combustion Synthesis of Binary/Ternary Co x Ni1− x Ferrite for Electrochemical Hydrogen and Oxygen Evolution. ACS Omega 2021, 6, 33024–33032. [Google Scholar] [CrossRef]
  11. Tang, Y.-J.; Liu, C.-H.; Huang, W.; Wang, X.-L.; Dong, L.-Z.; Li, S.-L.; Lan, Y.-Q. Bimetallic carbides-based nanocomposite as superior electrocatalyst for oxygen evolution reaction. ACS Appl. Mater. Interfaces 2017, 9, 16977–16985. [Google Scholar] [CrossRef] [PubMed]
  12. Xu, Y.; Xie, L.; Li, D.; Yang, R.; Jiang, D.; Chen, M. Engineering Ni(OH)2 nanosheet on CoMoO4 nanoplate array as efficient electrocatalyst for oxygen evolution reaction. ACS Sustain. Chem. Eng. 2018, 6, 16086–16095. [Google Scholar] [CrossRef]
  13. Yang, Y.; Lin, Z.; Gao, S.; Su, J.; Lun, Z.; Xia, G.; Chen, J.; Zhang, R.; Chen, Q. Tuning electronic structures of nonprecious ternary alloys encapsulated in graphene layers for optimizing overall water splitting activity. ACS Catal. 2017, 7, 469–479. [Google Scholar] [CrossRef]
  14. Wen, Y.; Li, C.-X.; Wang, J.; Qian, Y.-Y.; Shen, F.-C. Hierarchically Self-Supporting Phosphorus-Doped CoMoO4 Nanoflowers Arrays toward Efficient Hydrogen Evolution Reaction. ACS Appl. Energy Mater. 2022, 5, 6814–6822. [Google Scholar] [CrossRef]
  15. Suryanto, B.H.; Wang, Y.; Hocking, R.K.; Adamson, W.; Zhao, C. Overall electrochemical splitting of water at the heterogeneous interface of nickel and iron oxide. Nat. Commun. 2019, 10, 5599. [Google Scholar] [CrossRef]
  16. Liu, Z.-Z.; Shang, X.; Dong, B.; Chai, Y.-M. Triple Ni-Co-Mo metal sulfides with one-dimensional and hierarchical nanostructures towards highly efficient hydrogen evolution reaction. J. Catal. 2018, 361, 204–213. [Google Scholar] [CrossRef]
  17. Zhang, X.; Yu, X.; Zhang, L.; Zhou, F.; Liang, Y.; Wang, R. Molybdenum phosphide/carbon nanotube hybrids as pH-universal electrocatalysts for hydrogen evolution reaction. Adv. Funct. Mater. 2018, 28, 1706523. [Google Scholar] [CrossRef]
  18. 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]
  19. Lin, Y.; Liu, M.; Pan, Y.; Zhang, J. Porous Co–Mo phosphide nanotubes: An efficient electrocatalyst for hydrogen evolution. J. Mater. Sci. 2017, 52, 10406–10417. [Google Scholar] [CrossRef]
  20. Gong, Y.; Yang, Z.; Lin, Y.; Wang, J.; Pan, H.; Xu, Z. Hierarchical heterostructure NiCo2O4@CoMoO4/NF as an efficient bifunctional electrocatalyst for overall water splitting. J. Mater. Chem. A 2018, 6, 16950–16958. [Google Scholar] [CrossRef]
  21. Liu, Y.; Xu, P.; Ye, Z.; Cen, T.; Peng, X.; Yuan, D.; Wu, H. P doped CoMoO4/RGO as an efficient hybrid catalyst for hydrogen evolution. Int. J. Hydrogen Energy 2020, 45, 15157–15165. [Google Scholar] [CrossRef]
  22. Qian, J.; Wang, T.; Xia, B.; Xi, P.; Gao, D. Zn-doped MoSe2 nanosheets as high-performance electrocatalysts for hydrogen evolution reaction in acid media. Electrochim. Acta 2019, 296, 701–708. [Google Scholar] [CrossRef]
  23. Qu, Y.; Medina, H.; Wang, S.W.; Wang, Y.C.; Chen, C.W.; Su, T.Y.; Manikandan, A.; Wang, K.; Shih, Y.C.; Chang, J.W.; et al. Wafer scale phase-engineered 1T-and 2H-MoSe2/Mo core–shell 3D-hierarchical nanostructures toward efficient electrocatalytic hydrogen evolution reaction. Adv. Mater. 2016, 28, 9831–9838. [Google Scholar] [CrossRef]
  24. Ling, T.; Zhang, T.; Ge, B.; Han, L.; Zheng, L.; Lin, F.; Xu, Z.; Hu, W.B.; Du, X.W.; Davey, K.; et al. Well-dispersed nickel-and zinc-tailored electronic structure of a transition metal oxide for highly active alkaline hydrogen evolution reaction. Adv. Mater. 2019, 31, 1807771. [Google Scholar] [CrossRef]
  25. Chen, Z.; Kronawitter, C.X.; Koel, B.E. Facet-dependent activity and stability of Co3O4 nanocrystals towards the oxygen evolution reaction. Phys. Chem. Chem. Phys. 2015, 17, 29387–29393. [Google Scholar] [CrossRef]
  26. Ling, T.; Yan, D.-Y.; Wang, H.; Jiao, Y.; Hu, Z.; Zheng, Y.; Zheng, L.; Mao, J.; Liu, H.; Du, X.-W.; et al. Activating cobalt (II) oxide nanorods for efficient electrocatalysis by strain engineering. Nat. Commun. 2017, 8, 1509. [Google Scholar] [CrossRef] [PubMed]
  27. 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, 128, 5363–5367. [Google Scholar] [CrossRef]
  28. Zhang, H.; Zhang, J.; Li, Y.; Jiang, H.; Li, C. Continuous oxygen vacancy engineering of the Co3O4 layer for an enhanced alkaline electrocatalytic hydrogen evolution reaction. J. Mater. Chem. A 2019, 7, 13506–13510. [Google Scholar] [CrossRef]
  29. Ouyang, C.; Wang, X.; Wang, S. Phosphorus-doped CoS2 nanosheet arrays as ultra-efficient electrocatalysts for the hydrogen evolution reaction. Chem. Commun. 2015, 51, 14160–14163. [Google Scholar] [CrossRef]
  30. Jing, S.; Zhang, L.; Luo, L.; Lu, J.; Yin, S.; Shen, P.K.; Tsiakaras, P. N-doped porous molybdenum carbide nanobelts as efficient catalysts for hydrogen evolution reaction. Appl. Catal. B Environ. 2018, 224, 533–540. [Google Scholar] [CrossRef]
  31. Ren, M.; Guo, X.; Huang, S. Transition metal atoms (Fe, Co, Ni, Cu, Zn) doped RuIr surface for the hydrogen evolution reaction: A first-principles study. Appl. Surf. Sci. 2021, 556, 149801. [Google Scholar] [CrossRef]
  32. Jiao, F.; Li, J.; Wang, J.; Lin, Y.; Gong, Y.; Jing, X. Regulating the electronic structure of CoMoO4 microrod by phosphorus doping: An efficient electrocatalyst for the hydrogen evolution reaction. Dalton Trans. 2020, 49, 13152–13159. [Google Scholar] [CrossRef] [PubMed]
  33. Sharma, P.; Minakshi Sundaram, M.; Watcharatharapong, T.; Laird, D.; Euchner, H.; Ahuja, R. Zn metal atom doping on the surface plane of one-dimesional NiMoO4 nanorods with improved redox chemistry. ACS Appl. Mater. Interfaces 2020, 12, 44815–44829. [Google Scholar] [CrossRef]
  34. Costa, R.K.; Teles, S.C.; Siqueira, K.P. The relationship between crystal structures and thermochromism in CoMoO4. Chem. Pap. 2021, 75, 237–248. [Google Scholar] [CrossRef]
  35. Pirhashemi, M.; Habibi-Yangjeh, A. Facile fabrication of novel ZnO/CoMoO4 nanocomposites: Highly efficient visible-light-responsive photocatalysts in degradations of different contaminants. J. Photochem. Photobiol. A Chem. 2018, 363, 31–43. [Google Scholar] [CrossRef]
  36. Mandal, M.; Ghosh, D.; Giri, S.; Shakir, I.; Das, C.K. Polyaniline-wrapped 1D CoMoO4·0.75H2O nanorods as electrode materials for supercapacitor energy storage applications. RSC Adv. 2014, 4, 30832–30839. [Google Scholar] [CrossRef]
  37. Zhang, X.; Lou, S.; Zeng, Y. Facile fabrication of a novel visible light active g-C3N4-CoMoO4 heterojunction with largely improved photocatalytic performance. Mater. Lett. 2020, 281, 128661. [Google Scholar] [CrossRef]
  38. Rafieezadeh, M.; Kianfar, A.H. Fabrication of heterojunction ternary Fe3O4/TiO2/CoMoO4 as a magnetic photocatalyst for organic dyes degradation under sunlight irradiation. J. Photochem. Photobiol. A Chem. 2022, 423, 113596. [Google Scholar] [CrossRef]
  39. Vinothkumar, V.; Abinaya, M.; Chen, S.-M. Ultrasonic assisted preparation of CoMoO4 nanoparticles modified electrochemical sensor for chloramphenicol determination. J. Solid State Chem. 2021, 302, 122392. [Google Scholar] [CrossRef]
  40. Ahmed, J.; Ubaidullah, M.; Ahmad, T.; Alhokbany, N.; Alshehri, S.M. Synthesis of graphite oxide/cobalt molybdenum oxide hybrid nanosheets for enhanced electrochemical performance in supercapacitors and the oxygen evolution reaction. ChemElectroChem 2019, 6, 2524–2530. [Google Scholar] [CrossRef]
  41. Badawi, A.; Althobaiti, M.; Alharthi, S.S.; Alharbi, A.N.; Alkathiri, A.A.; Alomairy, S.E. Effect of zinc doping on the structure and optical properties of iron oxide nanostructured films prepared by spray pyrolysis technique. Appl. Phys. A 2022, 128, 123. [Google Scholar] [CrossRef]
  42. Kumar, L.; Kumar, P.; Kar, M. Cation distribution by Rietveld technique and magnetocrystalline anisotropy of Zn substituted nanocrystalline cobalt ferrite. J. Alloys Compd. 2013, 551, 72–81. [Google Scholar] [CrossRef]
  43. Fei, B.; Chen, Z.; Ha, Y.; Wang, R.; Yang, H.; Xu, H.; Wu, R. Anion-cation co-substitution activation of spinel CoMoO4 for efficient oxygen evolution reaction. Chem. Eng. J. 2020, 394, 124926. [Google Scholar] [CrossRef]
  44. Liu, Z.; Zhan, C.; Peng, L.; Cao, Y.; Chen, Y.; Ding, S.; Xiao, C.; Lai, X.; Li, J.; Wei, S. A CoMoO4–Co2Mo3O8 heterostructure with valence-rich molybdenum for a high-performance hydrogen evolution reaction in alkaline solution. J. Mater. Chem. A 2019, 7, 16761–16769. [Google Scholar] [CrossRef]
  45. Sadeghi, E.; Peighambardoust, N.S.; Aydemir, U. Tailoring the Morphology of Cost-Effective Vanadium Diboride Through Cobalt Substitution for Highly Efficient Alkaline Water Oxidation. Inorg. Chem. 2021, 60, 19457–19466. [Google Scholar] [CrossRef] [PubMed]
  46. Yu, N.; Zhang, Z.-J.; Yuan, X.-Q.; Liu, H.-J.; Zhou, Y.-L.; Lv, R.-Q.; Liu, Y.-B.; Chai, Y.-M.; Dong, B. Molten salt assisted Co1−xAgxMoO4 with lattice Ag doping and oxygen vacancy for stable water oxidation. Appl. Surf. Sci. 2023, 614, 156075. [Google Scholar] [CrossRef]
  47. Sun, H.; Tung, C.-W.; Qiu, Y.; Zhang, W.; Wang, Q.; Li, Z.; Tang, J.; Chen, H.-C.; Wang, C.; Chen, H.M. Atomic metal–support interaction enables reconstruction-free dual-site electrocatalyst. J. Am. Chem. Soc. 2021, 144, 1174–1186. [Google Scholar] [CrossRef]
  48. Liu, D.; Zhang, C.; Yu, Y.; Shi, Y.; Yu, Y.; Niu, Z.; Zhang, B. Hydrogen evolution activity enhancement by tuning the oxygen vacancies in self-supported mesoporous spinel oxide nanowire arrays. Nano Res. 2018, 11, 603–613. [Google Scholar] [CrossRef]
  49. Li, Y.; Chen, S.; Wu, X.; Zhang, H.; Zhang, J. A hybrid zeolitic imidazolate framework-derived ZnO/ZnMoO4 heterostructure for electrochemical hydrogen production. Dalton Trans. 2021, 50, 11365–11369. [Google Scholar] [CrossRef]
  50. Brijesh, K.; Bindu, K.; Shanbhag, D.; Nagaraja, H. Chemically prepared Polypyrrole/ZnWO4 nanocomposite electrodes for electrocatalytic water splitting. Int. J. Hydrogen Energy 2019, 44, 757–767. [Google Scholar] [CrossRef]
  51. Selvan, R.K.; Gedanken, A. The sonochemical synthesis and characterization of Cu1−xNixWO4 nanoparticles/nanorods and their application in electrocatalytic hydrogen evolution. Nanotechnology 2009, 20, 105602. [Google Scholar] [CrossRef] [PubMed]
  52. Padmanathan, N.; Shao, H.; Razeeb, K.M. Honeycomb micro/nano-architecture of stable β-NiMoO4 electrode/catalyst for sustainable energy storage and conversion devices. Int. J. Hydrogen Energy 2020, 45, 30911–30923. [Google Scholar] [CrossRef]
  53. Zhu, J.; Lv, W.; Yang, Y.; Huang, L.; Yu, W.; Wang, X.; Han, Q.; Dong, X. Hexagonal NiMoO4-MoS2 nanosheet heterostructure as a bifunctional electrocatalyst for urea oxidation assisted overall water electrolysis. New J. Chem. 2022, 46, 10280–10288. [Google Scholar] [CrossRef]
  54. Huang, S.; Meng, Y.; Cao, Y.; Yao, F.; He, Z.; Wang, X.; Pan, H.; Wu, M. Amorphous NiWO4 nanoparticles boosting the alkaline hydrogen evolution performance of Ni3S2 electrocatalysts. Appl. Catal. B Environ. 2020, 274, 119120. [Google Scholar] [CrossRef]
  55. Du, X.; Huang, C.; Zhang, X. Synthesis of CoMoO4/Co9S8 network arrays on nickel foam as efficient urea oxidation and hydrogen evolution catalyst. Int. J. Hydrogen Energy 2019, 44, 19595–19602. [Google Scholar] [CrossRef]
  56. Sun, H.; Xu, X.; Song, Y.; Zhou, W.; Shao, Z. Designing high-valence metal sites for electrochemical water splitting. Adv. Funct. Mater. 2021, 31, 2009779. [Google Scholar] [CrossRef]
  57. Sadeghi, E.; Peighambardoust, N.S.; Khatamian, M.; Unal, U.; Aydemir, U. Metal doped layered MgB2 nanoparticles as novel electrocatalysts for water splitting. Sci. Rep. 2021, 11, 3337. [Google Scholar] [CrossRef]
  58. Rani, B.J.; Swathi, S.; Yuvakkumar, R.; Ravi, G.; Kumar, P.; Babu, E.S.; Alfarraj, S.; Alharbi, S.A.; Velauthapillai, D. Solvothermal synthesis of CoMoO4 nanostructures for electrochemical applications. J. Mater. Sci. Mater. Electron. 2021, 32, 5989–6000. [Google Scholar] [CrossRef]
  59. Sun, H.; Li, J.-G.; Lv, L.; Li, Z.; Ao, X.; Xu, C.; Xue, X.; Hong, G.; Wang, C. Engineering hierarchical CoSe/NiFe layered-double-hydroxide nanoarrays as high efficient bifunctional electrocatalyst for overall water splitting. J. Power Source 2019, 425, 138–146. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic illustration of preparation procedure of Co1-xZnxMoO4, (b) XRD patterns of the prepared Co1-xZnxMoO4 samples, and (c) Crystal structure of CoMoO4.
Figure 1. (a) Schematic illustration of preparation procedure of Co1-xZnxMoO4, (b) XRD patterns of the prepared Co1-xZnxMoO4 samples, and (c) Crystal structure of CoMoO4.
Catalysts 13 00798 g001
Figure 2. Morphological, structural, and microstructural characterizations of the samples. SEM images of (a) CoMoO4, (b) Co0.9Zn0.1MoO4, (c) Co0.7Zn0.3MoO4, (d) Co0.5Zn0.5MoO4, and (e) Co0.3Zn0.7MoO4, respectively, (f) TEM photographs related to Co0.5Zn0.5MoO4 (inset: SAED pattern), white square showing the high-magnification TEM images of nanorods, green and red squares showing the HR-TEM images of lattice spacing of Co0.5Zn0.5MoO4, and (g) HAADF-STEM image and the corresponding EDS–STEM mappings of Co0.5Zn0.5MoO4.
Figure 2. Morphological, structural, and microstructural characterizations of the samples. SEM images of (a) CoMoO4, (b) Co0.9Zn0.1MoO4, (c) Co0.7Zn0.3MoO4, (d) Co0.5Zn0.5MoO4, and (e) Co0.3Zn0.7MoO4, respectively, (f) TEM photographs related to Co0.5Zn0.5MoO4 (inset: SAED pattern), white square showing the high-magnification TEM images of nanorods, green and red squares showing the HR-TEM images of lattice spacing of Co0.5Zn0.5MoO4, and (g) HAADF-STEM image and the corresponding EDS–STEM mappings of Co0.5Zn0.5MoO4.
Catalysts 13 00798 g002
Figure 3. Spectroscopic characterizations of the samples (a) From left to right XPS survey of CoMoO4 and Co0.5Zn0.5MoO4, XPS Co 2p, Mo 3d, and O 1s for CoMoO4, (b) From left to right, XPS Zn 2p, Co 2p, Mo 3d, and O 1s for Co0.5Zn0.5MoO4.
Figure 3. Spectroscopic characterizations of the samples (a) From left to right XPS survey of CoMoO4 and Co0.5Zn0.5MoO4, XPS Co 2p, Mo 3d, and O 1s for CoMoO4, (b) From left to right, XPS Zn 2p, Co 2p, Mo 3d, and O 1s for Co0.5Zn0.5MoO4.
Catalysts 13 00798 g003
Figure 4. Electrocatalytic hydrogen-evolution performances. (a) HER polarization curves of the samples, (b) 3D bar graphs of overpotentials at 10 and 50 mA cm−2, (c) HER Tafel plots obtained by polarization curves, (d) Capacitive current density versus scan rate curves for ECSA measurements, (e) EIS spectra recorded at 0 V versus RHE, (f) Long-term HER stability test of Co0.5Zn0.5MoO4 at an applied current density of 10 mA cm−2, (g) HER polarization curves of best-performing Co0.5Zn0.5MoO4 before and after long-term HER, and (h) HER polarization curves of Co0.5Zn0.5MoO4 before and after 1000 CV scans at 100 mV s−1.
Figure 4. Electrocatalytic hydrogen-evolution performances. (a) HER polarization curves of the samples, (b) 3D bar graphs of overpotentials at 10 and 50 mA cm−2, (c) HER Tafel plots obtained by polarization curves, (d) Capacitive current density versus scan rate curves for ECSA measurements, (e) EIS spectra recorded at 0 V versus RHE, (f) Long-term HER stability test of Co0.5Zn0.5MoO4 at an applied current density of 10 mA cm−2, (g) HER polarization curves of best-performing Co0.5Zn0.5MoO4 before and after long-term HER, and (h) HER polarization curves of Co0.5Zn0.5MoO4 before and after 1000 CV scans at 100 mV s−1.
Catalysts 13 00798 g004
Figure 5. Post-electrolysis characterizations of Co0.5Zn0.5MoO4 after long-term HER test. (a) SEM image (inset: low-magnification SEM image), (b) TEM and HR-TEM images (inset: SAED pattern), and (c) XPS Co 2p, Mo 3d, O 1s, and Zn 2p spectra.
Figure 5. Post-electrolysis characterizations of Co0.5Zn0.5MoO4 after long-term HER test. (a) SEM image (inset: low-magnification SEM image), (b) TEM and HR-TEM images (inset: SAED pattern), and (c) XPS Co 2p, Mo 3d, O 1s, and Zn 2p spectra.
Catalysts 13 00798 g005
Figure 6. (a) Schematic presentation of typical Volmer–Heyrovsky pathway for alkaline HER, (b) Amount of gas theoretically calculated and experimentally measured for HER using Co0.5Zn0.5MoO4 at 10 mA cm−2 (inset: homemade electrolysis cell), and (c) Quantification of Faradaic efficiency derived from calculated and measured H2.
Figure 6. (a) Schematic presentation of typical Volmer–Heyrovsky pathway for alkaline HER, (b) Amount of gas theoretically calculated and experimentally measured for HER using Co0.5Zn0.5MoO4 at 10 mA cm−2 (inset: homemade electrolysis cell), and (c) Quantification of Faradaic efficiency derived from calculated and measured H2.
Catalysts 13 00798 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

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. https://doi.org/10.3390/catal13050798

AMA Style

Chamani S, Sadeghi E, Unal U, Peighambardoust NS, Aydemir U. Tuning Electrochemical Hydrogen-Evolution Activity of CoMoO4 through Zn Incorporation. Catalysts. 2023; 13(5):798. https://doi.org/10.3390/catal13050798

Chicago/Turabian Style

Chamani, Sanaz, Ebrahim Sadeghi, Ugur Unal, Naeimeh Sadat Peighambardoust, and Umut Aydemir. 2023. "Tuning Electrochemical Hydrogen-Evolution Activity of CoMoO4 through Zn Incorporation" Catalysts 13, no. 5: 798. https://doi.org/10.3390/catal13050798

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