Next Article in Journal
Investigation into the Sonodynamic Activity of Three Newly Synthesized Derivatives of Ciprofloxacin
Previous Article in Journal
Antidepressant Sertraline Synergistically Enhances Paclitaxel Efficacy by Inducing Autophagy in Colorectal Cancer Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nitrogen-Tungsten Oxide Nanostructures on Nickel Foam as High Efficient Electrocatalysts for Benzyl Alcohol Oxidation

1
College of Material, Chemistry and Chemical Engineering, Hangzhou Normal University, Hangzhou 311121, China
2
Kharkiv Institute, Hangzhou Normal University, Hangzhou 311121, China
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(16), 3734; https://doi.org/10.3390/molecules29163734 (registering DOI)
Submission received: 27 May 2024 / Revised: 5 August 2024 / Accepted: 5 August 2024 / Published: 7 August 2024
(This article belongs to the Section Electrochemistry)

Abstract

:
Electrocatalytic alcohol oxidation (EAO) is an attractive alternative to the sluggish oxygen evolution reaction in electrochemical hydrogen evolution cells. However, the development of high-performance bifunctional electrocatalysts is a major challenge. Herein, we developed a nitrogen-doped bimetallic oxide electrocatalyst (WO-N/NF) by a one-step hydrothermal method for the selective electrooxidation of benzyl alcohol to benzoic acid in alkaline electrolytes. The WO-N/NF electrode features block-shaped particles on a rough, inhomogeneous surface with cracks and lumpy nodules, increasing active sites and enhancing electrolyte diffusion. The electrode demonstrates exceptional activity, stability, and selectivity, achieving efficient benzoic acid production while reducing the electrolysis voltage. A low onset potential of 1.38 V (vs. RHE) is achieved to reach a current density of 100 mA cm−2 in 1.0 M KOH electrolyte with only 0.2 mmol of metal precursors, which is 396 mV lower than that of water oxidation. The analysis reveals a yield, conversion, and selectivity of 98.41%, 99.66%, and 99.74%, respectively, with a Faradaic efficiency of 98.77%. This work provides insight into the rational design of a highly active and selective catalyst for electrocatalytic alcohol oxidation.

1. Introduction

In recent times, the depletion of fossil fuels and the pressing concern for environmental pollution have led to a surge in interest in generating renewable and clean energy sources [1,2,3,4]. Hydrogen energy is considered a promising energy source to substitute fossil fuels and provide a viable solution to the energy crisis and environmental issues [5,6]. Water electrolysis has a promising future for renewable energy in the hydrogen production industry, owing to its eco-friendly and sustainable characteristics [7,8]. However, there are still some limitations to large-scale hydrogen production via water electrolysis [9,10]. Firstly, the overpotentials required for the anodic oxygen evolution reaction (OER) and cathodic hydrogen evolution reaction (HER) throughout the water electrolysis procedure are excessive, consequently resulting in higher energy consumption [11,12]. In addition, the economic value of the anode by-product oxygen is low in the traditional approach of water electrolysis [13].
In light of the challenges associated with conventional water electrolysis, utilizing electrocatalytic organic oxidation reactions emerges as a more viable alternative [12,14]. Specifically, substituting the OER with the electrocatalytic oxidation of benzyl alcohol not only reduces the required potential for water electrolysis but also yields benzoic acid, a high-value chemical pivotal in both industrial production and fundamental research. This green approach, integrating electrocatalytic alcohol oxidation with hydrogen production, presents distinct advantages, including a lower potential, prevention of gas mixing, and the generation of value-added products. These merits position it as a promising research focus in the field [15,16]. Electrocatalytic oxidation technology leverages variations in electrode potential to drive electron transfer, expediting the synthesis of organic compounds. This approach avoids the use of additional toxic reagents, thereby reducing environmental pollution. It is also in line with the principles of sustainable development within modern industry and environmental conservation. Notably, the prevalent use of noble metals (e.g., Pt, Ir, Rh, Ru, etc.) or their oxides as electrode catalysts in electrochemistry faces limitations in commercial applications due to their prohibitively high costs [16].
As a three-dimensional (3D) porous substrate, nickel foam exhibits remarkable electrical conductivity, facilitating electron diffusion. Additionally, its high specific surface area promotes the even distribution of catalysts, and its porous structure enables the swift diffusion of gases that are produced. Xu and coworkers employed in situ construction of NiCo2O4 nanosheet catalysts on nickel foam for electrocatalytic benzyl alcohol oxidation reactions, which required a mere 1.46 V (vs. RHE) to achieve a current density of 100 mA·cm−2. After 2 h of reaction under optimized experimental conditions, the NiCo2O4/NF catalyst achieved over 95% conversion of benzyl alcohol and demonstrated high selectivity to benzoic acid, as well as high Faradaic efficiency and excellent stability [17]. Moreover, transition metal tungstates (TMTs) containing abundant element tungsten (W) are regarded as one of the most promising electrode materials for electrochemical energy storage compared to noble metal electrode materials [18]. Compared to traditional metals, tungsten (W) offers superior stability and extended service life in electrochemical applications, enhancing its appeal as an electrocatalyst. Tungsten-based materials also retain their performance under harsh electrochemical conditions, unlike other metals, which may degrade more quickly in strong oxidative environments [19,20]. TMTs manifest high durability, cost-effectiveness, and environmental compatibility [21]. However, TMTs face challenges in capacity and stability due to low electrical conductivity and nanoparticle aggregation [22]. Therefore, a common approach to address these issues involves synthesizing dispersed nanocomposites with enhanced electrical conductivity and a strategy supported by studies showing substantial improvement in electrocatalyst efficiency through nitrogen (N) doping [23,24,25]. This strategy not only enhances the material’s electrical conductivity but also fortifies the stability of nickel foam [26,27]. For instance, Wang and co-workers loaded a novel copper-nickel nitride (Cu1Ni2-N) which was enriched with a Cu4N/Ni3N interface on carbon fibers [28]. As a result, the electrode demonstrated a hydrogen precipitation overpotential of 71.4 mV at a current density of 10 mA·cm−2 and exhibited outstanding stability over 75 h. However, current research primarily focuses on improving the OER reaction activity, with limited exploration of TMT materials for electrocatalytic alcohol oxidation and hydrogen production.
In this work, a one-step hydrothermal method was employed to synthesize WO-N/NF using nickel foam and ammonium metatungstate. The resulting product was effectively used in the catalytic oxidation reaction of benzyl alcohol. The impact of varied solution concentrations and hydrothermal reaction times on the electrodes was explored. The microscopic morphology, crystal structure, elemental composition and content, and elemental valence states of the synthesized electrodes’ surface were comprehensively investigated with scanning electron microscopy, X-ray diffraction, energy dispersal spectroscopy, and X-ray photoelectron spectroscopy. The hydrogen precipitation performance of the electrode in alkaline electrolytes was characterized with the aid of an electrochemical workstation. The optimal hydrothermal reaction time and reaction solution concentration were considered during the analysis.

2. Results and Discussion

2.1. Characterization of WO-N/NF

WO-N nickel foam electrodes were synthesized via a one-step hydrothermal approach, utilizing nickel foam as the substrate and ammonium metatungstate as the nitrogen and tungsten sources. The hydrothermal reaction induced a discernible change in the surface color of the Ni foam, transitioning from silver to yellowish green (Figure S1), which implies a new substance was produced. As the SEM images showed (Figure 1a, Figure S2), the coating grows directly on the nickel foam substrate, and its surface shows inhomogeneity and roughness accompanied by the presence of cracks. The formation of these cracks may be associated with the presence of residual stresses within the coating [29]. In addition, lumpy nodules appeared on the surface of the electrode, increasing the specific surface area of the electrode. This structure helps provide more active sites and increases the contact area between the electrolyte and the electrode surface, facilitating electrolyte diffusion. As can be seen from the higher magnification images (Figure 1b), the surface of the WO-N/NF electrode is characterized by block-shaped particles. It is observed that these particles are grown on the coating with a distributed pattern, which may suggest a staggered arrangement of nanosheets [30]. This direct-grown block coating avoids the use of adhesives, thereby reducing the distance of diffusive electron transport and lowering the electron transfer resistance, thereby facilitating the oxidation reaction. TEM imaging of the powders obtained from the Ni foam substrate further validated the morphology of the sheet arrays (Figure 1c). The high-resolution TEM image (Figure 1d) reveals the presence of three lattice fringes with a spacing of 0.365 nm, 0.372 nm, and 0.176 nm, which can be attributed to the (220) and (020) lattice planes of WO3 and the (200) lattice plane of Ni, respectively. The corresponding SAED pattern (Figure 1d) shows well-defined spots, which can be readily indexed to monoclinic WO3 [19]. In addition, the elemental mapping image (Figure 2) reveals a uniform distribution of Ni, W, O, and N on the electrode surface. To determine the elemental composition and content of the WO-N/NF electrode’s surface, EDS analysis was conducted. The EDS spectrum in Figure S3 and Table S1 provides information on the composition of the WO-N/NF electrode, allowing for the determination of the presence and content of each element.
XPS was utilized to analyze the surface composition and elemental chemistries of the WO-N/NF electrode. Figure S4 confirmed the presence of Ni, W, O, and N, consistent with the EDS results. The high-resolution spectrum of Ni 2p (Figure 3a) shows double peaks at binding energies of 855.51 eV and 873.25 eV, indicating the presence of nickel in the Ni2+ state. Notably, two satellite peaks at 861.39 eV and 879.23 eV are observed [31,32,33]. A minor peak at 851.79 eV corresponds to the 2p3/2 orbital of Ni0 from the nickel foam substrate, indicating that certain fragments on the nickel foam surface have oxidized to Ni2+. Higher valence Ni effectively reduces the reaction potential required for its oxidation [34]. As depicted in Figure S5, there are double peaks with binding energies of 855.23 eV and 872.97 eV, indicating the presence of nickel in the Ni2+ state. Compared with hydrothermal NF, it is noteworthy that this double peak also appears in the WO-N/NF sample after W doping, with the corresponding binding energy shifted by 0.28 eV to the higher energy region. This result indicates the successful doping of W elements and shows that the doping affects the electronic structure of the catalyst material. It has been reported that metal nitrides are easily oxidized during the oxygen evolution reaction to produce the highly reactive species NiOOH [35,36]. Therefore, the doping of W causes the Ni in the WO-N/NF samples to lose electrons, creating an electron-deficient state that is more likely to produce more reactive species (NiOOH), thus promoting the activity of the hydrogen evolution reaction. The high-resolution energy spectrum of W 4f (Figure 3b) exhibited two primary peaks at binding energies of 36.84 eV and 34.72 eV, attributed to W4f5/2 and W4f7/2, respectively, indicating the presence of W6+. In addition, the small peak on the left with a binding energy of 40.63 eV represents W 5p3/2. According to the literature, W6+ is predominantly present in catalysts such as WO3 or WO24− [37,38,39,40]. In the O 1s high-resolution energy spectrum (Figure 3c), the peak at 532.01 eV corresponds to water molecules adsorbed on the electrode surface, while the peak at 530.7 eV is attributed to the Ni/W-O bond [29]. The N 1s spectrum (Figure 3d) exhibits two peaks at 398.40 eV and 400.00 eV, corresponding to the Ni/W-N and -NH bonds, respectively [41,42,43]. The above experimental results show that the one-step hydrothermal method successfully prepared nickel foam electrodes doped with tungsten and nitrogen elements.
XRD analysis was conducted to probe the phase state of the sample. As shown in Figure 4a, the electrode exhibits three major diffraction peaks at 44.70°, 52.02°, and 76.64°, corresponding to the (111), (200), and (220) crystal planes of metallic Ni (JCPDS No. 65-2865). The thin thickness of the hydrothermal accumulation layer may contribute to the full display of the nickel foam peaks within the matrix. In Figure 4b, the local magnification of the electrode from 40° to 50° reveals a leftward shift of the WO-N/NF electrode relative to the standard card on the Ni (111) surface. According to the Bragg equation, this shift suggests the solid dissolution of W and N atoms in the Ni lattice, precluding the appearance of W- and N-related diffraction peaks [44,45].

2.2. Electrocatalytic Performance

To better investigate the catalytic oxidation activity of the WO-N/NF electrode, a standard three-electrode system with a platinum (Pt) electrode as the counter electrode and Ag/AgCl as the reference electrode was used in 1 M KOH alkaline medium using an electrochemical workstation at a scan rate of 40 mV s−1.
Initially, to determine the appropriate amount of doping in WO-N/NF, we conducted Linear Sweep Voltammetry (LSV) to assess its electrocatalytic performance across different concentrations. Figure 5a illustrates the LSV curves for nickel foam electrodes with varied W-doped concentrations in 40 mL of 1.0 M KOH with 0.1 M benzyl alcohol electrolyte. Notably, the 0.2-WO-N/NF displays the most pronounced slope, indicative of accelerated current density growth, minimal alterations in overpotential, and superior electrocatalytic performance. Consequently, we designate 0.2-WO-N/NF as the optimal doping level. After this, a comprehensive analysis of the electrochemical properties of the WO-N/NF electrodes was conducted, encompassing cyclic voltammetry, linear scanning voltammetry, and electrochemical impedance spectroscopy.
For comparison, the catalytic activities of WO-N/NF, WO/NF, N/NF, hydrothermal NF, and bare NF were also measured under the same conditions. As depicted in the polarization curves (Figure 5b), the slope of the WO-N/NF electrode exceeds that of the other electrodes in an alkaline electrolyte containing benzyl alcohol, indicating superior catalytic oxidation performance [46,47]. Furthermore, the LSV curve of the WO-N/NF electrode was measured with and without 0.1 M benzyl alcohol (Figure S6). The electrode requires 1.52 V (vs. RHE) to achieve a current density of 20 mA cm−2 without benzyl alcohol, outperforming the benchmark electrolyzer consisting of a Pt-C/NF cathode and a RuO2/NF anode, which requires 1.60 V to reach the same current density [48]. The Tafel slope, a crucial parameter for evaluating electrode catalytic performance, was determined using the Tafel equation η = b l o g j + a , where a and b denote the intercept and Tafel slope, respectively. Figure 5c depicts the Tafel plots of the corresponding polarization curves. In this work, the slope for WO-N/NF is 33.6 mV dec−1, which is significantly lower than that of WO/NF (52.1 mV dec−1), N/NF (73.1 mV dec−1), hydrothermal NF (86.2 mV dec−1), and NF (144 mV dec−1), indicating a preferable HER kinetics on WO-N/NF. Generally, a smaller Tafel slope indicates a greater ability of the electrode material to transfer charge and a faster reaction rate of the electrode [25]. Thus, the WO-N/NF electrode exhibits a faster oxidation reaction rate.
Electrochemical Impedance Spectroscopy (EIS) tests were conducted to investigate the internal resistance and charge transfer at the surface of samples. Figure 5d illustrates the EIS profiles of the WO-N/NF and other electrodes under bias. All electrodes display a semicircular shape in their impedance graphs, with the WO-N/NF electrode exhibiting a markedly smaller radius than the others. Additionally, the equivalent circuit model used for the EIS measurements is shown, consisting of three elements: Rs, Rct, and CPE. These represent the solution resistance of the alkaline electrolyte, the charge transfer resistance at the electrode–electrolyte interface, and the constant phase element of the bilayer at the same interface, respectively. Zview software (Version 3.1) was used to fit the EIS data using this equivalent circuit. The fitting results, as detailed in Table 1, indicate that the charge transfer resistance of WO-N/NF is 1.36 Ω, which is lower than that of WO/NF (6.84 Ω), N/NF (8.3 Ω), hydrothermal NF (97.21 Ω), and NF (590.02 Ω). This suggests that ion diffusion in the electrolyte is faster and that the HER kinetic process is more rapid on WO-N/NF. Generally, a smaller charge transfer resistance translates to lower electron resistance during the transfer process, facilitating faster electron transfer and oxidation rates, ultimately leading to improved electrode oxidation performance [49]. The WO-N/NF electrode exhibits a faster oxidation reaction rate and better oxidation catalytic performance due to its significantly smaller charge transfer resistance compared to the other electrodes [50]. According to the Engel–Brewer bonding theory, when the d orbitals of the transition metal element W have an available electron density in excess of the number of d orbitals, alloying with a transition metal such as Ni occurs in the 3d84s2 configuration [51,52]. The synergistic interaction between these species contributes to enhancing the performance of the electrode. The incorporation of the W element improved the charge transfer resistance of the Ni electrode and enhanced its oxidation performance. This improvement can be attributed to the electron transfer caused by the introduction of W, which improves the electronic structure of the electrode and enhances its catalytic performance.
The electrochemical active surface area (ECSA) reflects the number of active sites on the electrode’s surface [53]. A greater ECSA implies an increased exposure of active sites, correlating with the heightened catalytic activity of the electrode [53,54]. In Figure 5e, WO-N/NF shows a larger Cdl (3.52 mF cm−2) compared to WO/NF (3.41 mF cm−2), N/NF (2.62 mF cm−2), hydrothermal NF (2.89 mF cm−2), and pure NF (2.61 mF cm−2). This implies that WO-N/NF contains the most active sites, thereby promoting the HER performance. The ECSA of the electrode can be calculated using the double-layer capacitance, and the Equation (1) for ECSA is as follows:
E C S A = C d l C s
The stability and durability of the electrode material are important parameters for evaluating the oxidation performance of the catalyst. In industrial applications, the electrode material must maintain the stability of the morphology and composition of the electrode surface, as well as long-term durability over a prolonged period of operation [54]. Cyclic voltammetry (CV) scanning, a commonly employed technique, accelerates the degradation of electrode materials. Material stability is determined by comparing anodic polarization curves before and after multiple cycles [55]. Figure 5f displays the anodic polarization curves of the WO-N/NF electrode before and after 2000 CV cycles in a 1 M KOH solution at room temperature. The polarization curves after 2000 CV cycles almost coincide with the original polarization curves, indicating that the WO-N/NF electrode exhibits good oxidation performance. Further tests involved chronoamperometry, conducted five times consecutively for four hours at a potential of 0.77 V (vs. Ag/AgCl) (Figure S7). This potential was chosen as it ensures no significant anodic oxygen evolution reaction. The figure demonstrates that the electrolysis curve of the WO-N/NF electrode maintains stability in 1M KOH electrolyte, indicating good stability. The gradual decrease in electrode current density with electrolysis time, stabilizing after five repeated uses, indicates sustained performance over time.
In order to define the products in the process of electrocatalytic oxidation of benzyl alcohol, electrochemical oxidation tests were systematically conducted at a constant potential. Chronoamperometry experiments were repetitively performed over five consecutive sessions, each lasting four hours, at a stable potential of 1.87 V, ensuring negligible occurrence of the OER. The concentrations of benzyl alcohol and its oxidation by-products during the electrolysis process were meticulously tracked using HPLC. Figure 6a shows the high-performance liquid chromatogram of benzyl alcohol during its electrocatalytic oxidation. There are three distinct peaks at different positions corresponding to different electrocatalytic products, which can correspond to benzyl alcohol (peak 1), benzaldehyde (peak 2), and benzoic acid (peak 3), in that order. Meanwhile, a significant amount of hydrogen was discharged from the cathodic electrolytic cell during the electrocatalytic oxidation of benzyl alcohol, verifying that the electrocatalytic oxidation of benzyl alcohol in the anodic electrolytic cell was enhanced. As depicted in Figure 6b, the concentration of benzyl alcohol gradually diminishes throughout the electrochemical oxidation reaction. By the 6000s mark, benzyl alcohol oxidation attains near completion, leading to the quantitative characterization of the acidified anode product. The analysis reveals an impressive yield, conversion, and selectivity of 98.41%, 99.66%, and 99.74%, respectively, coupled with a commendable Faradaic efficiency of 98.77%.
The reusability of the WO-NF catalyst for benzyl alcohol oxidation was investigated. As shown in Figure 6c, after five cycles, the yield, conversion, and selectivity persist at impressive levels of 97.28%, 98.63%, and 98.13%, respectively, with the Faradaic efficiency maintaining a commendable 98.55% (Figure 6d), demonstrating sustained high performance without notable reduction. After five consecutive cycles, no significant changes in the morphology and structure of WO-N/Ni were observed (Figures S8 and S9), indicating its high stability. The Ni 2p and W 4f spectrums (Figure S10) reveal that Ni2+/Ni3+ and W6+ remain the dominant species in WO-N/Ni. Furthermore, the relative contents of metal-O and metal-N do not show significant changes after the reaction. Furthermore, the oxidation reaction follows a first-order reaction, as shown by the fitting lines of −ln(C/C0) vs. time in Figure S11 [56,57]. Here, C and t represent the concentration of benzyl alcohol and the oxidation time, respectively. The oxidation rate constants for the WO-N/NF and NF catalysts are 2.11 × 10−4 s−1 and 1.40 × 10−4 s−1, respectively. In addition, Table S2 presents a comparative assessment of the catalytic performance of WO-N/NF against previously reported electrodes. These results suggest that the WO-N/NF catalyst exhibits superior catalytic activity in the oxidation of benzyl alcohol compared to the bare NF substrate.

3. Experimental Section

3.1. Chemicals

Nickel foam was purchased from Kunshan Guangjiayuan New Materials Co., Ltd. (Kunshan, China). Ammonium metatungstate (NH4)6H2W12O40·XH2O was purchased from Shanghai Aladdin Biochemical Technology Co., Ltd. (Shanghai, China). Benzyl alcohol (99%), benzoic acid (99%), and benzaldehyde (99%) were purchased from Energy Chemical (Shanghai, China). Acetone (AR), hydrochloric acid (AR), and acetic acid (AR) were purchased from Sinopharm Chemical Reagent (Shanghai, China). Methanol (HPLC) was purchased from Sinopharm Chemical Reagent and Sigma-Aldrich (St. Louis, MO, USA). All chemical reagents were used without further purification.

3.2. Characterization

The crystal structure of the materials was evaluated by X-ray diffraction (XRD, Bruker D8, Bruker, Billarica, MA, USA). The apparent morphologies of the materials and EDS images of the material were investigated using a scanning electron microscope (SEM, Sigma 500, Oberkochen, Germany). Transmission electron microscope (HRTEM, Hillsboro, OR, USA, 200 kV) images were taken using an FEI Tecnai F20 microscope (FEI, Hillsboro, OR, USA). X-ray photoelectron spectroscopy (XPS, Waltham, MA, USA) analysis of the surface electronic states was performed using a Thermo ESCALAB 250Xi (Waltham, WA, USA).

3.3. Synthesis of WO-N/NF

The materials were prepared using a one-step hydrothermal method (Figure S12). A 10 × 40 × 1 mm3 nickel foam was initially intercepted and then ultrasonically washed in sequence using acetone, 1 M hydrochloric acid, and deionized water for 30 min. Subsequently, the foam was vacuum-dried. A total of 0.2 mmol ammonium metatungstate was dissolved in 40 mL deionized water by stirring at room temperature using a stirring table. The configured solution and nickel foam were moved to a 100 mL hydrothermal reactor and heated in an oven at 160 °C for 8 h then cooled to room temperature; the nickel foam with color change was collected, rinsed with deionized water to remove the impurities on the surface, and dried at 50 °C in a vacuum oven for 12 h to obtain the WO-N/NF. The doping content of W was adjusted by varying the molar amount of ammonium metatungstate. By adding x mmol of ammonium metatungstate, the product obtained was defined as x-WO-N/NF.

3.4. Electrochemical Measurement

Electrochemical measurements were carried out using an electrochemical workstation (Autolab PGSTAT128N, Metrohm AG, Herisau, Switzerland). The electrocatalytic oxidation reaction of benzyl alcohol was conducted in an H-type cell double electrolyzer separated by a Nafion-117 proton exchange membrane (Chemours, Wilmington, DE, USA) placed between the two electrolyzers. A three-electrode system was employed in the reaction, utilizing WO-N/NF as the working electrode, a counter electrode composed of a platinum foil measuring 1 cm × 1 cm, and a reference electrode of silver chloride (Ag/AgCl). The measured potentials were converted to reversible hydrogen electrodes using the following equation, Equation (2):
E v s . R H E = E + E v s . A g / A g C l + 0.059 × p H
The electrolyte utilized for testing electrochemical hydrogen evolution reaction performance was a 40 mL 1 M KOH solution.

3.5. Benzyl Alcohol Electrocatalytic Oxidation Performance Test

This study examined the electrocatalytic oxidation of benzyl alcohol utilizing WO-N/NF as the electrocatalyst. Benzyl alcohol and its catalytic products were quantified through an Agilent 1260 high-performance liquid chromatography (HPLC) (Santa Clara, CA, USA) equipped with a UV detector. The corresponding conversion of benzyl alcohol and its product selectivity were subsequently calculated. The specific detection parameters are as follows: the liquid chromatography column was C18; the mobile phases were methanol and 10% aqueous acetic acid (the volume ratio was 4:6) at a flow rate of 0.6 mL/min; the temperature of the detection column was 40 °C; and the detection wavelength was 245 nm. The concentration of the components was quantified using the external standard method, with the peak area integral as the vertical coordinate and the concentration as the horizontal coordinate. Subsequently, the standard curve was fitted and graphed as depicted in Figure S13. The results showed that the retention times of benzyl alcohol, benzaldehyde, and benzoic acid in the mixed control solution were approximately 10 min, 14 min, and 15 min, respectively. In addition, the benzyl alcohol conversion, product selectivity, and yield were calculated using Equations (3)–(5):
B e n z y l   a l c o h o l   c o n v e r s i o n   % = n r e a c t e d   b e n z y l   a l c o h o l n i n i t i a l   b e n z y l   a l c o h o l × 100 %
B e n z o i c   a c i d   s e l e c t i v i t y   % = n b e n z o i c   a c i d   p r o d u c e d n r e a c t e d   b e n z y l   a l c o h o l × 100 %
B e n z o i c   a c i d   y i e l d   % = n b e n z o i c   a c i d   p r o d u c e d n i n i t i a l   b e n z y l   a l c o h o l × 100 %
The formula for Faradaic efficiency (FE) is given by Equation (6):
F E = n z F Q
where n is the actual number of moles of benzoic acid produced, z is the number of electrons transferred (4 in this reaction), F is Faraday’s constant (96,485 C/mol), and Q is the total charge passed through the circuit.

4. Conclusions

In summary, the N-doped bimetallic oxide electrocatalyst on Ni foam (WO-N/NF) was successfully fabricated via a one-step hydrothermal method. Owing to its optimized composition, structure, and morphology, it shows a high HER activity with potential of 1.38 V (vs. RHE) to reach a current density of 100 mA cm−2. The doping of tungsten induces electron transfer in WO-N/NF samples, enhancing NiOOH formation and catalytic activity for the hydrogen evolution reaction. The incorporation of nitrogen atoms into ammonium metatungstate effectively modulates the electronic structure of transition metal atoms, enhancing the catalytic performance. Furthermore, the WO-N/NF was applied to the preparation of benzoic acid by electrocatalytic oxidation of benzyl alcohol and obtained high conversion, yield, and Faradaic efficiency, reaching 98.41%, 99.66%, and 98.77%, respectively. This approach opens up new possibilities for electrocatalytic organic reactions, emphasizing the potential of our electrocatalyst in advancing green and sustainable technologies in the field.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29163734/s1, Figure S1: Optical photograph of bare NF and WO-N/NF; Figure S2: SEM images of (a,b) WO/NF, (c,d) N/NF, (e,f) Hydrothermal NF; Figure S3: EDS spectrum of WO-N/NF electrode; Figure S4: XPS survey spectrum of WO-N/NF electrode; Figure S5: XPS spectrums of Ni 2p, O 1s for the obtained hydrothermal NF electrode; Figure S6: LSV curves of WO-N/NF under OER and BAO; Figure S7: Chronoamperometry tests of electrocatalytic benzyl alcohol oxidation for five cycles; Figure S8: SEM image of WO-N/NF electrode after 5 cycles of measurement; Figure S9: XRD pattern of WO-N/NF electrode after 5 cycles of measurement; Figure S10: XPS spectrums of Ni 2p (a), W 4f (b), O 1s (c), and N 1s (d) for the obtained WO-N/NF electrode after 5 cycles of measurement; Figure S11: First-order kinetics models of the electrochemical oxidation of BAO for WO-N/NF and NF at 1.87 V vs. RHE; Figure S12: Scheme for the synthesis of WO-N/NF electrocatalyst for benzyl alcohol oxidation; Figure S13: HPLC chromatogram of the standard mixed solution at different concentrations (a) benzyl alcohol, (c) benzaldehyde and (e) benzoic acid. The standard curves of (b) benzyl alcohol, (d) benzaldehyde and (f) benzoic acid; Table S1: Composition and content of each element in WO-N/NF electrode; Table S2: Comparison of the catalytic performance of the WO-N/NF and the stand-of-the-art catalysts towards benzyl alcohol electrochemical oxidation; Table S3: Each abbreviation corresponds to the full name; Synthesis of WO/NF; Synthesis of N/NF; Synthesis of hydrothermal NF. References [15,17,58,59,60,61,62,63,64,65,66,67] are cited in Supplementary Materials file.

Author Contributions

Y.Z. (Yizhen Zhu), K.C. and W.X. contributed to the design of the study. Y.Z. (Yizhen Zhu) and X.C. conducted experiments and data analysis. K.C. and W.X. provided technical advice and result interpretation. Y.Z. (Yizhen Zhu) and Y.Z. (Yuanyao Zhang) conducted catalyst characterization and corresponding data processing. Y.Z. (Yizhen Zhu), Z.Z. and H.C. conducted catalysis performance test and corresponding data processing. Y.Z. (Yizhen Zhu), K.C. and W.X. wrote the manuscript and the ESI material; all authors commented on and amended both documents. All authors discussed and contributed to the work. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Sci-Tech Project of Zhejiang (LGG21B020001), the Natural Science Foundation of China (22178078), Sci-Tech Project of Hangzhou (20201203B137), the National Student Innovation and Entrepreneurship Programme (202410346041), and the Zhejiang Province Science and Technology Innovation Activity Programme for College Students.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in [Nitrogen-Tungsten Oxide Nanostructures on Nickel Foam as High Efficient Electrocatalysts for Benzyl Alcohol Oxidation].

Acknowledgments

We gratefully acknowledge the financial support by the Sci-Tech Project of Zhejiang (LGG21B020001), the Natural Science Foundation of China (22178078), Sci-Tech Project of Hangzhou (20201203B137), the National Student Innovation and Entrepreneurship Programme (202410346041), and the Zhejiang Province Science and Technology Innovation Activity Programme for College Students.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. van Putten, R.-J.; van der Waal, J.C.; de Jong, E.; Rasrendra, C.B.; Heeres, H.J.; de Vries Hydroxymethylfurfural, J.G. A Versatile Platform Chemical Made from Renewable Resources. Chem. Rev. 2013, 113, 1499–1597. [Google Scholar] [CrossRef]
  2. Hauer, A. Energy Storage Solutions for Future Energy Systems. In Advances in Energy Storage; John Wiley & Sons Ltd.: Hoboken, NJ, USA, 2022; pp. 1–14. [Google Scholar] [CrossRef]
  3. Kabeyi, M.J.B.; Olanrewaju, O.A. Sustainable Energy Transition for Renewable and Low Carbon Grid Electricity Generation and Supply. Front. Energy Res. 2022, 9, 743114. [Google Scholar] [CrossRef]
  4. Li, J.; Liu, H.; An, Z.; Kong, Y.; Huang, L.; Duan, D.; Long, R.; Yang, P.; Jiang, Y.-Y.; Liu, J.; et al. Nitrogen-doped carbon for selective pseudo-metal-free hydrodeoxygenation of 5-hydroxymethylfurfural to 2,5-dimethylfuran: Importance of trace iron impurity. J. Catal. 2023, 417, 396–407. [Google Scholar] [CrossRef]
  5. Zhang, P.; Sun, L. Electrocatalytic Hydrogenation and Oxidation in Aqueous Conditions. Chin. J. Chem. 2020, 38, 996–1004. [Google Scholar] [CrossRef]
  6. Li, C.-Q.; Yi, S.-S.; Liu, Y.; Niu, Z.-L.; Yue, X.-Z.; Liu, Z.-Y. In-situ constructing S-scheme/Schottky junction and oxygen vacancy on SrTiO3 to steer charge transfer for boosted photocatalytic H2 evolution. Chem. Eng. J. 2021, 417, 129231. [Google Scholar] [CrossRef]
  7. Nishioka, S.; Osterloh, F.E.; Wang, X.; Mallouk, T.E.; Maeda, K. Photocatalytic water splitting. Nat. Rev. Methods Primers 2023, 3, 42. [Google Scholar] [CrossRef]
  8. Tan, C.; Geng, S.; Tan, Z.; Wang, G.; Pu, L.; Guo, X. Integrated energy system–Hydrogen natural gas hybrid energy storage system optimization model based on cooperative game under carbon neutrality. J. Energy Storage 2021, 38, 102539. [Google Scholar] [CrossRef]
  9. Capurso, T.; Stefanizzi, M.; Torresi, M.; Camporeale, S.M. Perspective of the role of hydrogen in the 21st century energy transition. Energy Convers. Manag. 2022, 251, 114898. [Google Scholar] [CrossRef]
  10. Abdin, Z.; Zafaranloo, A.; Rafiee, A.; Mérida, W.; Lipiński, W.; Khalilpour, K.R. Hydrogen as an energy vector. Renew. Sustain. Energy Rev. 2020, 120, 109620. [Google Scholar] [CrossRef]
  11. Suen, N.-T.; Hung, S.-F.; Quan, Q.; Zhang, N.; Xu, Y.-J.; Chen, H.M. Electrocatalysis for the oxygen evolution reaction: Recent development and future perspectives. Chem. Soc. Rev. 2017, 46, 337–365. [Google Scholar] [CrossRef] [PubMed]
  12. Li, T.; Wang, Q.; Wu, J.; Sui, Y.; Tang, P.; Liu, H.; Zhang, W.; Li, H.; Wang, Y.; Cabot, A.; et al. Strain and Shell Thickness Engineering in Pd3Pb@Pt Bifunctional Electrocatalyst for Ethanol Upgrading Coupled with Hydrogen Production. Small 2023, 20, 2306178. [Google Scholar] [CrossRef] [PubMed]
  13. Miller, H.A.; Lavacchi, A.; Vizza, F. Storage of renewable energy in fuels and chemicals through electrochemical reforming of bioalcohols. Curr. Opin. Electrochem. 2020, 21, 140–145. [Google Scholar] [CrossRef]
  14. Tang, W.; Liu, Z.; Tan, Q.; Zhang, Y. Analysis of electric energy substitution potential and research on calculation method of carbon reduction under “carbon neutrality” background. IOP Conf. Ser. Earth Environ. Sci. 2022, 1087, 012026. [Google Scholar] [CrossRef]
  15. Ming, L.; Wu, X.-Y.; Wang, S.-S.; Wu, W.; Lu, C.-Z. Facile growth of transition metal hydroxide nanosheets on porous nickel foam for efficient electrooxidation of benzyl alcohol. Green Chem. 2021, 23, 7825–7830. [Google Scholar] [CrossRef]
  16. Llorente, M.J.; Nguyen, B.H.; Kubiak, C.P.; Moeller, K.D. Paired Electrolysis in the Simultaneous Production of Synthetic Intermediates and Substrates. J. Am. Chem. Soc. 2016, 138, 15110–15113. [Google Scholar] [CrossRef] [PubMed]
  17. Xu, M.; Geng, J.; Xu, H.; Zhang, S.; Zhang, H. In situ construction of NiCo2O4 nanosheets on nickel foam for efficient electrocatalytic oxidation of benzyl alcohol. Inorg. Chem. Front. 2023, 10, 2053–2059. [Google Scholar] [CrossRef]
  18. AlShehri, S.M.; Ahmed, J.; Ahamad, T.; Arunachalam, P.; Ahmad, T.; Khan, A. Bifunctional electro-catalytic performances of CoWO4 nanocubes for water redox reactions (OER/ORR). RSC Adv. 2017, 7, 45615–45623. [Google Scholar] [CrossRef]
  19. Su, C.Y.; Lin, C.K.; Yang, T.K.; Lin, H.-C.; Pan, C.T. Oxygen partial pressure effect on the preparation of nanocrystalline tungsten oxide powders by a plasma arc gas condensation technique. Int. J. Refract. Met. Hard Mater. 2008, 26, 423–428. [Google Scholar] [CrossRef]
  20. Yang, L.; Zhu, X.; Xiong, S.; Wu, X.; Shan, Y.; Chu, P.K. Synergistic WO3·2H2O Nanoplates/WS2 Hybrid Catalysts for High-Efficiency Hydrogen Evolution. ACS Appl. Mater. Interfaces 2016, 8, 13966–13972. [Google Scholar] [CrossRef] [PubMed]
  21. Kubacka, A.; Si, R.; Michorczyk, P.; Martínez-Arias, A.; Xu, W.; Hanson, J.C.; Rodriguez, J.A.; Fernández-García, M. Tungsten as an interface agent leading to highly active and stable copper–ceria water gas shift catalyst. Appl. Catal. B Environ. 2013, 132-133, 423–432. [Google Scholar] [CrossRef]
  22. Yu, P.; Wang, L.; Liu, X.; Fu, H.-G.; Yu, H.-T. CoWO4 nanopaticles wrapped by RGO as high capacity anode material for lithium ion batteries. Rare Met. 2017, 36, 411–417. [Google Scholar] [CrossRef]
  23. Wang, Y.; Zhang, M.; Liu, Y.; Zheng, Z.; Liu, B.; Chen, M.; Guan, G.; Yan, K. Recent Advances on Transition-Metal-Based Layered Double Hydroxides Nanosheets for Electrocatalytic Energy Conversion. Adv. Sci. 2023, 10, 2207519. [Google Scholar] [CrossRef] [PubMed]
  24. Theerthagiri, J.; Lee, S.J.; Murthy, A.P.; Madhavan, J.; Choi, M.Y. Fundamental aspects and recent advances in transition metal nitrides as electrocatalysts for hydrogen evolution reaction: A review. Curr. Opin. Solid State Mater. Sci. 2020, 24, 100805. [Google Scholar] [CrossRef]
  25. Wu, L.; Qin, H.; Ji, Z.; Zhou, H.; Shen, X.; Zhu, G.; Yuan, A. Nitrogen-Doped Carbon Dots Modified Fe–Co Sulfide Nanosheets as High-Efficiency Electrocatalysts toward Oxygen Evolution Reaction. Small 2023, 20, 2305965. [Google Scholar] [CrossRef] [PubMed]
  26. Ni, W.; Meibom, J.L.; Hassan, N.U.; Chang, M.; Chu, Y.-C.; Krammer, A.; Sun, S.; Zheng, Y.; Bai, L.; Ma, W.; et al. Synergistic interactions between PtRu catalyst and nitrogen-doped carbon support boost hydrogen oxidation. Nat. Catal. 2023, 6, 773–783. [Google Scholar] [CrossRef]
  27. Yuan, C.; Yang, L.; Hou, L.; Shen, L.; Zhang, X.; Lou, X.W. Growth of ultrathin mesoporous Co3O4 nanosheet arrays on Ni foam for high-performance electrochemical capacitors. Energy Environ. Sci. 2012, 5, 7883–7887. [Google Scholar] [CrossRef]
  28. Wang, Z.; Xu, L.; Huang, F.; Qu, L.; Li, J.; Owusu, K.A.; Liu, Z.; Lin, Z.; Xiang, B.; Liu, X.; et al. Copper–Nickel Nitride Nanosheets as Efficient Bifunctional Catalysts for Hydrazine-Assisted Electrolytic Hydrogen Production. Adv. Energy Mater. 2019, 9, 1900390. [Google Scholar] [CrossRef]
  29. Li, J.-K.; Wang, A.; Dong, X.-Y.; Huang, S.; Meng, Y.; Song, J.-L. Construction of 2D C,N-co-doped ZnO/Co3O4 over Ni(OH)2 mesoporous ultrathin nanosheets on Ni foam as high-performance electrocatalysts for benzyl-alcohol oxidation and accelerating hydrogen evolution. New J. Chem. 2023, 47, 5970–5976. [Google Scholar] [CrossRef]
  30. Yang, C.; Wang, C.; Zhou, L.; Duan, W.; Song, Y.; Zhang, F.; Zhen, Y.; Zhang, J.; Bao, W.; Lu, Y.; et al. Refining d-band center in Ni0.85Se by Mo doping: A strategy for boosting hydrogen generation via coupling electrocatalytic oxidation 5-hydroxymethylfurfural. Chem. Eng. J. 2021, 422, 130125. [Google Scholar] [CrossRef]
  31. Zhuang, Z.; Giles, S.A.; Zheng, J.; Jenness, G.R.; Caratzoulas, S.; Vlachos, D.G.; Yan, Y. Nickel supported on nitrogen-doped carbon nanotubes as hydrogen oxidation reaction catalyst in alkaline electrolyte. Nat. Commun. 2016, 7, 10141. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, Y.; Li, Y.; Ding, L.; Chen, Z.; Ong, A.; Lu, W.; Herng, T.S.; Li, X.; Ding, J. NiFe (sulfur)oxyhydroxide porous nanoclusters/Ni foam composite electrode drives a large-current-density oxygen evolution reaction with an ultra-low overpotential. J. Mater. Chem. A 2019, 7, 18816–18822. [Google Scholar] [CrossRef]
  33. Zhang, J.; Xu, Q.; Wang, J.; Hu, Y.; Jiang, H.; Li, C. Heterointerface engineered NiFe(OH)x/Ni3S2 electrocatalysts to overcome the scaling relationship for ultrahigh-current-density water oxidation. Sci. China Mater. 2023, 66, 634–640. [Google Scholar] [CrossRef]
  34. Ge, Z.; Fu, B.; Zhao, J.; Li, X.; Ma, B.; Chen, Y. A review of the electrocatalysts on hydrogen evolution reaction with an emphasis on Fe, Co and Ni-based phosphides. J. Mater. Sci. 2020, 55, 14081–14104. [Google Scholar] [CrossRef]
  35. Qin, Q.; Chen, L.; Wei, T.; Liu, X. MoS2/NiS Yolk–Shell Microsphere-Based Electrodes for Overall Water Splitting and Asymmetric Supercapacitor. Small 2019, 15, 1803639. [Google Scholar] [CrossRef] [PubMed]
  36. Dai, Z.; Geng, H.; Wang, J.; Luo, Y.; Li, B.; Zong, Y.; Yang, J.; Guo, Y.; Zheng, Y.; Wang, X.; et al. Hexagonal-Phase Cobalt Monophosphosulfide for Highly Efficient Overall Water Splitting. ACS Nano 2017, 11, 11031–11040. [Google Scholar] [CrossRef] [PubMed]
  37. Su, Y.; Zhang, H.; Liang, P.; Liu, K.; Cai, M.; Huang, Z.; Wang, C.-A.; Zhong, M. A new binder-free and conductive-additive-free TiO2/WO3-W integrative anode material produced by laser ablation. J. Power Sources 2018, 378, 362–368. [Google Scholar] [CrossRef]
  38. Du, Y.; Wang, W.; Zhao, H.; Liu, Y.; Li, S.; Wang, L. The rational doping of P and W in multi-stage catalysts to trigger Pt-like electrocatalytic performance. J. Mater. Chem. A 2020, 8, 25165–25172. [Google Scholar] [CrossRef]
  39. Wang, H.; Liu, T.; Bao, K.; Cao, J.; Feng, J.; Qi, J. W doping dominated NiO/NiS2 interfaced nanosheets for highly efficient overall water splitting. J. Colloid Interface Sci. 2020, 562, 363–369. [Google Scholar] [CrossRef] [PubMed]
  40. Wang, F.; Niu, S.; Liang, X.; Wang, G.; Chen, M. Phosphorus incorporation activates the basal plane of tungsten disulfide for efficient hydrogen evolution catalysis. Nano Res. 2022, 15, 2855–2861. [Google Scholar] [CrossRef]
  41. Cao, B.; Veith, G.M.; Diaz, R.E.; Liu, J.; Stach, E.A.; Adzic, R.R.; Khalifah, P.G. Cobalt Molybdenum Oxynitrides: Synthesis, Structural Characterization, and Catalytic Activity for the Oxygen Reduction Reaction. Angew. Chem. Int. Ed. 2013, 52, 10753–10757. [Google Scholar] [CrossRef] [PubMed]
  42. Wu, Y.; Liu, X.; Han, D.; Song, X.; Shi, L.; Song, Y.; Niu, S.; Xie, Y.; Cai, J.; Wu, S.; et al. Electron density modulation of NiCo2S4 nanowires by nitrogen incorporation for highly efficient hydrogen evolution catalysis. Nat. Commun. 2018, 9, 1425. [Google Scholar] [CrossRef] [PubMed]
  43. Li, Z.; Chen, Y.; Ji, S.; Tang, Y.; Chen, W.; Li, A.; Zhao, J.; Xiong, Y.; Wu, Y.; Gong, Y.; et al. Iridium single-atom catalyst on nitrogen-doped carbon for formic acid oxidation synthesized using a general host–guest strategy. Nat. Chem. 2020, 12, 764–772. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, T.; Wang, X.; Liu, Y.; Zheng, J.; Li, X. A highly efficient and stable biphasic nanocrystalline Ni–Mo–N catalyst for hydrogen evolution in both acidic and alkaline electrolytes. Nano Energy 2016, 22, 111–119. [Google Scholar] [CrossRef]
  45. Yang, M.; Li, J.; Ke, G.; Liu, B.; Dong, F.; Yang, L.; He, H.; Zhou, Y. WO3 homojunction photoanode: Integrating the advantages of WO3 different facets for efficient water oxidation. J. Energy Chem. 2021, 56, 37–45. [Google Scholar] [CrossRef]
  46. Voiry, D.; Chhowalla, M.; Gogotsi, Y.; Kotov, N.A.; Li, Y.; Penner, R.M.; Schaak, R.E.; Weiss, P.S. Best Practices for Reporting Electrocatalytic Performance of Nanomaterials. ACS Nano 2018, 12, 9635–9638. [Google Scholar] [CrossRef] [PubMed]
  47. Li, L.; Wang, P.; Shao, Q.; Huang, X. Metallic nanostructures with low dimensionality for electrochemical water splitting. Chem. Soc. Rev. 2020, 49, 3072–3106. [Google Scholar] [CrossRef] [PubMed]
  48. Liu, H.; Ma, X.; Hu, H.; Pan, Y.; Zhao, W.; Liu, J.; Zhao, X.; Wang, J.; Yang, Z.; Zhao, Q.; et al. Robust NiCoP/CoP Heterostructures for Highly Efficient Hydrogen Evolution Electrocatalysis in Alkaline Solution. ACS Appl. Mater. Interfaces 2019, 11, 15528–15536. [Google Scholar] [CrossRef] [PubMed]
  49. Wei, M.; Sun, Y.; Zhang, J.; Ai, F.; Xi, S.; Wang, J. High-entropy alloy nanocrystal assembled by nanosheets with d–d electron interaction for hydrogen evolution reaction. Energy Environ. Sci. 2023, 16, 4009–4019. [Google Scholar] [CrossRef]
  50. Zhao, Y.; Björk, E.M.; Yan, Y.; Schaaf, P.; Wang, D. Recent progress in transition metal based catalysts and mechanism analysis for alcohol electrooxidation reactions. Green Chem. 2024, 26, 4987–5003. [Google Scholar] [CrossRef]
  51. He, G.; Han, X.; Moss, B.; Weng, Z.; Gadipelli, S.; Lai, F.; Kafizas, A.G.; Brett, D.J.L.; Guo, Z.X.; Wang, H.; et al. Solid solution nitride/carbon nanotube hybrids enhance electrocatalysis of oxygen in zinc-air batteries. Energy Storage Mater. 2018, 15, 380–387. [Google Scholar] [CrossRef]
  52. Wang, H.; Carter, E.A. Metal-metal bonding in Engel-Brewer intermetallics: “anomalous” charge transfer in zirconium-platinum (ZrPt3). J. Am. Chem. Soc. 1993, 115, 2357–2362. [Google Scholar] [CrossRef]
  53. Wang, Z.; Goddard, W.A.; Xiao, H. Potential-dependent transition of reaction mechanisms for oxygen evolution on layered double hydroxides. Nat. Commun. 2023, 14, 4228. [Google Scholar] [CrossRef] [PubMed]
  54. Zhou, W.; Wu, X.-J.; Cao, X.; Huang, X.; Tan, C.; Tian, J.; Liu, H.; Wang, J.; Zhang, H. Ni3S2 nanorods/Ni foam composite electrode with low overpotential for electrocatalytic oxygen evolution. Energy Environ. Sci. 2013, 6, 2921–2924. [Google Scholar] [CrossRef]
  55. Guo, T.; Li, L.; Wang, Z. Recent Development and Future Perspectives of Amorphous Transition Metal-Based Electrocatalysts for Oxygen Evolution Reaction. Adv. Energy Mater. 2022, 12, 2200827. [Google Scholar] [CrossRef]
  56. Chai, K.; Yang, X.; Shen, R.; Chen, J.; Su, W.; Su, A. A high activity mesoporous Pt@KIT-6 nanocomposite for selective hydrogenation of halogenated nitroarenes in a continuous-flow microreactor. Nanoscale Adv. 2023, 5, 5649–5660. [Google Scholar] [CrossRef] [PubMed]
  57. Hu, K.; Zhang, M.; Liu, B.; Yang, Z.; Li, R.; Yan, K. Efficient electrochemical oxidation of 5-hydroxymethylfurfural to 2,5-furandicarboxylic acid using the facilely synthesized 3D porous WO3/Ni electrode. Mol. Catal. 2021, 504, 111459. [Google Scholar] [CrossRef]
  58. Zheng, J.; Chen, X.; Zhong, X.; Li, S.; Liu, T.; Zhuang, G.; Li, X.; Deng, S.; Mei, D.; Wang, J.-G. Hierarchical Porous NC@CuCo Nitride Nanosheet Networks: Highly Efficient Bifunctional Electrocatalyst for Overall Water Splitting and Selective Electrooxidation of Benzyl Alcohol. Adv. Funct. Mater. 2017, 27, 1704169. [Google Scholar] [CrossRef]
  59. Cao, Y.; Zhang, D.; Kong, X.; Zhang, F.; Lei, X. Multi-vacancy Co3O4 on nickel foam synthesized via a one-step hydrothermal method for high-efficiency electrocatalytic benzyl alcohol oxidation. J. Mater. Sci. 2021, 56, 6689–6703. [Google Scholar] [CrossRef]
  60. Huang, H.; Yu, C.; Han, X.; Huang, H.; Wei, Q.; Guo, W.; Wang, Z.; Qiu, J. Ni, Co hydroxide triggers electrocatalytic production of high-purity benzoic acid over 400 mA cm−2. Energy Environ. Sci. 2020, 13, 4990–4999. [Google Scholar] [CrossRef]
  61. Yin, Z.; Zheng, Y.; Wang, H.; Li, J.; Zhu, Q.; Wang, Y.; Ma, N.; Hu, G.; He, B.; Knop-Gericke, A.; et al. Engineering Interface with One-Dimensional Co3O4 Nanostructure in Catalytic Membrane Electrode: Toward an Advanced Electrocatalyst for Alcohol Oxidation. ACS Nano 2017, 11, 12365–12377. [Google Scholar] [CrossRef] [PubMed]
  62. Chen, X.; Zhong, X.; Yuan, B.; Li, S.; Gu, Y.; Zhang, Q.; Zhuang, G.; Li, X.; Deng, S.; Wang, J.-G. Defect engineering of nickel hydroxide nanosheets by Ostwald ripening for enhanced selective electrocatalytic alcohol oxidation. Green Chem. 2019, 21, 578–588. [Google Scholar] [CrossRef]
  63. You, B.; Liu, X.; Liu, X.; Sun, Y. Efficient H2 Evolution Coupled with Oxidative Refining of Alcohols via A Hierarchically Porous Nickel Bifunctional Electrocatalyst. ACS Catal. 2017, 7, 4564–4570. [Google Scholar] [CrossRef]
  64. Cui, X.; Chen, M.; Xiong, R.; Sun, J.; Liu, X.; Geng, B. Ultrastable and efficient H2 production via membrane-free hybrid water electrolysis over a bifunctional catalyst of hierarchical Mo–Ni alloy nanoparticles. J. Mater. Chem. A 2019, 7, 16501–16507. [Google Scholar] [CrossRef]
  65. Huang, Y.; Yang, R.; Anandhababu, G.; Xie, J.; Lv, J.; Zhao, X.; Wang, X.; Wu, M.; Li, Q.; Wang, Y. Cobalt/Iron(Oxides) Heterostructures for Efficient Oxygen Evolution and Benzyl Alcohol Oxidation Reactions. ACS Energy Lett. 2018, 3, 1854–1860. [Google Scholar] [CrossRef]
  66. Song, Y.; Yuan, M.; Su, W.; Guo, D.; Chen, X.; Sun, G.; Zhang, W. Ultrathin Two-Dimensional Bimetal–Organic Framework Nanosheets as High-Performance Electrocatalysts for Benzyl Alcohol Oxidation. Inorg. Chem. 2022, 61, 7308–7317. [Google Scholar] [CrossRef] [PubMed]
  67. Zhang, C.; Ci, S.; Peng, X.; Huang, J.; Cai, P.; Ding, Y.; Wen, Z. Tri-profit electrolysis for energy-efficient production of benzoic acid and H2. J. Energy Chem. 2021, 54, 30–35. [Google Scholar] [CrossRef]
Figure 1. Surface morphology characterization of WO-N/NF electrode: (a,b) SEM; (c) TEM, and (d) HRTEM image and SAED pattern (inset in (d)).
Figure 1. Surface morphology characterization of WO-N/NF electrode: (a,b) SEM; (c) TEM, and (d) HRTEM image and SAED pattern (inset in (d)).
Molecules 29 03734 g001
Figure 2. EDS mapping images of the WO-N/NF nanocomposite.
Figure 2. EDS mapping images of the WO-N/NF nanocomposite.
Molecules 29 03734 g002
Figure 3. XPS spectrums of Ni 2p (a), W 4f (b), O 1s (c), and N 1s (d) for the obtained WO-N/NF electrode.
Figure 3. XPS spectrums of Ni 2p (a), W 4f (b), O 1s (c), and N 1s (d) for the obtained WO-N/NF electrode.
Molecules 29 03734 g003
Figure 4. Structure test diagram of electrodes: (a) XRD patterns of WO-N/NF; (b) localized magnified XRD patterns at 40° to 50°.
Figure 4. Structure test diagram of electrodes: (a) XRD patterns of WO-N/NF; (b) localized magnified XRD patterns at 40° to 50°.
Molecules 29 03734 g004
Figure 5. (a) LSV plots for WO-N/NF with different concentrations of precursors. (b) LSV curves of WO-N/NF, WO/NF, N/NF, hydrothermal NF and NF at the anode and (c) Tafel plots derived from (b); (d) Nyquist diagrams (the inset is an equivalent circuit model); (e) double layer capacitance diagram; (f) anodic polarization curves before and after 2000 CV.
Figure 5. (a) LSV plots for WO-N/NF with different concentrations of precursors. (b) LSV curves of WO-N/NF, WO/NF, N/NF, hydrothermal NF and NF at the anode and (c) Tafel plots derived from (b); (d) Nyquist diagrams (the inset is an equivalent circuit model); (e) double layer capacitance diagram; (f) anodic polarization curves before and after 2000 CV.
Molecules 29 03734 g005
Figure 6. (a) HPLC traces of catalysis by WO-N/NF at an actual potential of 0.77 V (vs. Ag/AgCl) in 40 mL of 1.0 M KOH with 0.1 M benzyl alcohol; (b) the concentration of benzyl alcohol and its oxidated products during the electrolysis; (c) the conversion of benzyl alcohol, and the selectivity and yield of benzoic acid; (d) the corresponding faradaic efficiencies for seven electrolysis cycles.
Figure 6. (a) HPLC traces of catalysis by WO-N/NF at an actual potential of 0.77 V (vs. Ag/AgCl) in 40 mL of 1.0 M KOH with 0.1 M benzyl alcohol; (b) the concentration of benzyl alcohol and its oxidated products during the electrolysis; (c) the conversion of benzyl alcohol, and the selectivity and yield of benzoic acid; (d) the corresponding faradaic efficiencies for seven electrolysis cycles.
Molecules 29 03734 g006
Table 1. Equivalent circuit model element parameters of NF and WO-N/NF.
Table 1. Equivalent circuit model element parameters of NF and WO-N/NF.
ElectrodeRs/(Ω)Rct/(Ω)
NF0.30590.02
WO-N/NF0.321.36
WO/NF0.326.84
N/NF0.3388.52
Hydrothermal NF0.3397.21
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

Zhu, Y.; Chen, X.; Zhang, Y.; Zhu, Z.; Chen, H.; Chai, K.; Xu, W. Nitrogen-Tungsten Oxide Nanostructures on Nickel Foam as High Efficient Electrocatalysts for Benzyl Alcohol Oxidation. Molecules 2024, 29, 3734. https://doi.org/10.3390/molecules29163734

AMA Style

Zhu Y, Chen X, Zhang Y, Zhu Z, Chen H, Chai K, Xu W. Nitrogen-Tungsten Oxide Nanostructures on Nickel Foam as High Efficient Electrocatalysts for Benzyl Alcohol Oxidation. Molecules. 2024; 29(16):3734. https://doi.org/10.3390/molecules29163734

Chicago/Turabian Style

Zhu, Yizhen, Xiangyu Chen, Yuanyao Zhang, Zhifei Zhu, Handan Chen, Kejie Chai, and Weiming Xu. 2024. "Nitrogen-Tungsten Oxide Nanostructures on Nickel Foam as High Efficient Electrocatalysts for Benzyl Alcohol Oxidation" Molecules 29, no. 16: 3734. https://doi.org/10.3390/molecules29163734

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop