Next Article in Journal
Pt3Mn/SiO2 + ZSM-5 Bifunctional Catalyst for Ethane Dehydroaromatization
Next Article in Special Issue
Expediting Corrosion Engineering for Sulfur-Doped, Self-Supporting Ni-Fe Layered Dihydroxide in Efficient Aqueous Oxygen Evolution
Previous Article in Journal
Non-Oxidative Coupling of Methane Catalyzed by Heterogeneous Catalysts Containing Singly Dispersed Metal Sites
Previous Article in Special Issue
Nano-Sheets of CsNiVF6 Pyrochlore Electrocatalyst for Enhanced Urea Oxidation and Hydrogen Green Production Reactions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Modification of NiSe2 Nanoparticles by ZIF-8-Derived NC for Boosting H2O2 Production from Electrochemical Oxygen Reduction in Acidic Media

Key Laboratory for Green Chemical Technology of Ministry of Education, Collaborative Innovation Center of Chemical Science and Engineering, School of Chemical Engineering and Technology, Tianjin University, Tianjin 300072, China
*
Author to whom correspondence should be addressed.
Submission received: 11 May 2024 / Revised: 27 May 2024 / Accepted: 31 May 2024 / Published: 3 June 2024
(This article belongs to the Special Issue Electrocatalysis for Hydrogen/Oxygen Evolution Reactions)

Abstract

:
The two-electron oxygen reduction reaction (2e ORR) has emerged as an attractive alternative for H2O2 production. Developing efficient earth-abundant transition metal electrocatalysts and reaction mechanism exploration for H2O2 production are important but remain challenging. Herein, a nitrogen-doped carbon-coated NiSe2 (NiSe2@NC) electrocatalyst was prepared by successive annealing treatment. Benefiting from the synergistic effect between the NiSe2 nanoparticles and NC, the 2e ORR activity, selectivity, and stability of NiSe2@NC in 0.1 M HClO4 was greatly enhanced, with the yield of H2O2 being 4.4 times that of the bare NiSe2 nanoparticles. The in situ Raman spectra and density functional theory (DFT) calculation revealed that the presence of NC was beneficial for regulating the electronic state of NiSe2 and optimizing the adsorption free energy of *OOH, which could enhance the adsorption of O2, stabilize the O-O bond, and boost the production of H2O2. This work provides an effective strategy to improve the performance of the transition metal chalcogenide for 2e ORR to H2O2.

1. Introduction

Hydrogen peroxide (H2O2) has been a high-value-added chemical with wide applications [1,2,3]. The industrial production of H2O2 uses an anthraquinone process, which has many disadvantages such as high energy consumption and pollution [4,5,6]. In contrast, the preparation of H2O2 using electrocatalytic oxygen reduction can be carried out at room temperature and pressure. The environment-friendly electrosynthesis of H2O2 through 2e ORR is regarded as a promising alternative [7,8]. As is generally known, ORR is an electrocatalytic reaction proceeding with a four-electron (4e) pathway or two-electron (2e) pathway when applying different electrocatalysts. The 4e ORR reaction has been intensively studied and applied in battery conversion systems such as fuel cells or metal–air batteries [9,10,11], while the development of efficient electrocatalysts for the selective 2e ORR pathway toward H2O2 remains challenging at present [1,9].
To date, various catalysts have been developed for electrocatalytic oxygen reduction to produce H2O2, but most of them showed a significantly improved 2e ORR performance only under alkaline media [12,13,14,15]. However, from the perspective of practical application, H2O2 is prone to decomposition in the alkaline environment for a long time, which greatly limits its application in actual industry and daily life [16,17,18]. On the contrary, acidic media can not only effectively stabilize H2O2, but it can also be more suitable for application scenarios such as electro-Fenton, showing more attractive advantages [7,19,20,21]. Therefore, researchers tend to develop high-performance electrocatalysts under acidic media. Precious metal-based catalysts have been approved to show the best 2e ORR performance under acidic media, but are limited by their high cost and scarcity [22,23,24]. Therefore, transition metal chalcogenides have attracted more attention because of their advantages such as their low cost, abundant reserves, high stability, and easy adjustment of electronic structure [25,26,27].
In particular, transition metal selenides (TMSs) have been regarded as a promising 2e ORR catalyst under acidic media [28,29,30]. NiSe2 is a typical TMS catalyst, but most reports focus on the field of water splitting or energy storage in previous studies [31,32]. Recently, NiSe2 has also been confirmed to be effective for 2e ORR reaction [33]. Due to the large atomic radius, high electronegativity of the Se element, and metal-like property, NiSe2 can effectively weaken the interactions of the outer metal electrons and then optimize the activation of oxygen-containing intermediates to promote the generation of H2O2. However, NiSe2 often exhibits poor electrical conductivity and tends to aggregate into larger particles, which could affect the electron transfer rate [34,35]. It is well-known that the carbon materials have a high electrical conductivity and stability; in the meantime, they could stabilize the nanoparticles as the supports or coating agents [36,37]. Accordingly, compositing NiSe2 with highly conductive carbon materials could make use of both advantages to facilitate the formation of H2O2 from 2e ORR in acidic media. In addition, metal–organic frameworks (MOFs) have the advantages of large porosity, large surface area, and tunable morphology and structure, and they are often used as an effective precursor for preparing efficient electrocatalysts [38]. Therefore, using MOFs precursors could provide a good platform for the effective combination of NiSe2 with a carbon material.
Based on the above considerations, in order to enhance the 2e ORR performance of NiSe2, a nitrogen-doped carbon-coated NiSe2 (NiSe2@NC) electrocatalyst was fabricated using ZIF-8 as the N-doped carbon precursor. The results showed that the introduction of NC could promote NiSe2 to form small-particle-size nanoparticles. In comparing with the bare NiSe2, NiSe2@NC could achieve a significantly improved H2O2 Faraday efficiency of 73% to 81%, H2O2 yield of 12 ppm to 53 ppm in 2 h if electrolysis, and long-term stability of 10 h in an H-type cell. In situ Raman spectroscopy and DFT calculation revealed that NiSe2@NC was beneficial for adjusting the electronic state of NiSe2 and optimizing the adsorption free energy of *OOH, which was conducive to promoting 2e ORR to produce H2O2. This work hopefully provides a valuable reference for the development of more efficient transition metal electrocatalysts toward 2e ORR.

2. Results and Discussion

2.1. Material Characterizations

The synthetic procedure of NiSe2@NC is displayed in Figure 1, which is prepared by the impregnation of Ni2+ with ZIF-8 and the successive annealing treatment, respectively. Firstly, Ni2+ ions were introduced into the self-assembled ZIF-8 to form Ni-ZIF-8. Then, the annealing treatment at 1000 °C under an Ar atmosphere was applied to form Ni-N-C. The subsequent pyrolysis process with selenium powder at 500 °C was performed to obtain the NiSe2@NC catalyst.
The morphologies of the catalysts were characterized by scanning electron microscopy (SEM) and transmission electron microscopy (TEM). As shown in Figure S1a, the ZIF-8 samples exhibit a rhombic dodecahedral morphology with the diameter of 200~250 nm. Figure 2a displays the N-doped carbon (NC) samples with a pristine morphology by the annealing treatment of ZIF-8 at 1000 °C under an Ar atmosphere, which have a smaller size (100~200 nm) than ZIF-8. And there is no formation of metal nanoparticles inside (Figure 2d), implying that the Zn NPs formed by annealing are completely removed under high-temperature calcination. The as-prepared NiSe2@NC samples (Figure 2b,e) show a disordered structure and a rough surface containing small nanoparticles (10~20 nm), indicating the successful composite of metal and NC. In contrast, pure NiSe2 particles (Figure 2c,f) show much larger aggregated particles in the absence of carbon carriers. Furthermore, the high-resolution TEM image (HRTEM) in Figure 2g shows that the lattice fringe of the nanoparticles is 0.266 nm, which corresponds to the (210) plane of NiSe2 [33], again validating the successful combination of NiSe2 and NC. Moreover, the elemental mapping images of NiSe2@NC shown in Figure 2h–k further indicate the presence of C, Ni, Se, and N elements and uniform distribution. These results imply that N-doped carbon-coated NiSe2 catalysts were prepared successfully. Meanwhile, the NiSe2 nanoparticles with a small particle size in NiSe2@NC could be beneficial to expose more reactive active sites and thus obtain a higher ORR activity.
The composition and the crystal structure of the samples were characterized by X-ray diffraction (XRD). The characteristic diffraction peaks of ZIF-8 (Figure S1b) are consistent with previous reports [39]. As shown in Figure 3a, NC has wide diffraction peaks at 24.8° and 43.6°, which can be indexed to the (002) and (100) planes of carbon, respectively. Both NiSe2@NC and NiSe2 show peaks at 29.8°, 33.5°, 36.8°, 42.8°, 50.6°, 55.4°, 57.8°, 62.2°, and 72.6°, corresponding to the (200), (210), (211), (220), (311), (230), (321), (400), and (421) planes of NiSe2 (PDF#65-1843) [40], respectively, which further confirms the successful synthesis of the NiSe2@NC catalyst. Combined with the Scherrer equation, the average crystal size of NiSe2 nanoparticles in NiSe2@NC could be calculated to be 20.2 nm, which is consistent with the TEM characterization result. In addition, the carbon properties were characterized by Raman spectroscopy as shown in Figure 3b. Both NC and NiSe2@NC have two characteristic peaks at 1350 cm−1 and 1590 cm−1, corresponding to the D and G bands of carbon, respectively. The D band represents disordered or defective carbon and the G band represents graphitized carbon [41]. The ID/IG values of NC and NiSe2@NC are 0.86 and 0.90, respectively, indicating the presence of defects in both catalysts, which have been reported to favor ORR activity [42].
For comparison, the other two NiSe2@NC samples (NiSe2@NC-L and NiSe2@NC-H) with lower and higher Ni addition were prepared under the same condition, respectively. Their morphologies and compositions were also observed. As shown in Figure S2a,b, the morphology of NiSe2@NC-H is similar to that of NiSe2@NC, while a large number of tubular structures exist in NiSe2@NC-L. The XRD pattern of NiSe2@NC-L (Figure S2c) shows a series of characteristic diffraction peaks of NiSe2 (PDF#65-1843) and a sharp diffraction peak at 26°, which is attributed to the (002) plane of graphitized carbon, corresponding to the formed carbon nanotubes [43]. In the case of increasing the Ni2+ addition, the diffraction peaks of NiSe2 disappear in NiSe2@NC-H, and new characteristic peaks appear at 33.0°, 44.6°, 50.5°, 60.2°, 61.6°, and 69.6°, which can be well indexed to Ni0.85Se (PDF#18-0888) [44]. These results indicate that the amount of Ni precursor addition could affect the formation of nickel selenide, and the NiSe2 phase could gradually transform into the Ni0.85Se phase with the increase in Ni addition.
The composition and valence state of NiSe2@NC were further determined by X-ray photoelectron spectroscopy (XPS). The signal peaks of C 1s, N 1s, Ni 2p, and Se 3d can be observed in Figure S3, confirming the existence of the four elements. The C 1s spectrum (Figure 4a) can be divided into two peaks ascribed to C-C (284.8 eV) and C-N (285.7 eV) [45]. The N 1s spectrum can be divided into three peaks at 398.2 eV, 400.0 eV, and 401.8 eV, corresponding to pyridinic N, pyrrolic N, and graphitic N, respectively (Figure 4b) [46]. The Ni 2p spectrum in Figure 4c shows the peak signals of Ni 2p3/2 and Ni 2p1/2. The fitting peaks at the binding energy of 853.4 eV and 870.9 eV are related to Ni2+, and the peaks at 855.6 eV and 873.8 eV are ascribed to Ni3+ (oxidation state). And the other two peaks at 860.1 eV and 879.1 eV can be assigned to the satellite peaks [47]. The Se 3d spectrum in Figure 4d has two peaks at 54.8 eV and 55.6 eV, which correspond to Se 3d5/2 and Se 3d3/2, indicating the existence form of Se22− [48]. And the peak at 59.0 eV can be assigned to oxidized Se (SeOx) [37,48].

2.2. Electrocatalytic Performance

The 2e ORR performance of NC, NiSe2@NC, and NiSe2 were evaluated by the RRDE system in O2-saturated 0.1 M HClO4. According to the polarization curves shown in Figure 5a, NC has the most positive onset potential (0.57 V vs. RHE), followed by NiSe2@NC (0.55 V vs. RHE) and NiSe2 (0.54 V vs. RHE). The disk current densities also present the same changing trend, which means NC has the highest ORR reaction kinetics among the three catalysts. Obviously, NiSe2@NC also shows a better ORR activity than bare NiSe2 due to the combination with N-doped carbon. Figure 5b shows the ring current densities of the three catalysts at different potentials. Compared with the other two catalysts, NiSe2@NC has the highest ring current density, which is about 1.2 and 1.8 times that of NC and NiSe2 at 0 V vs. RHE, respectively. This trend means that more O2 is converted to H2O2, which can be confirmed by the H2O2 selectivity and electron transfer number in Figure 5c,d. NiSe2@NC has the highest H2O2 selectivity of 77% and the lowest electron transfer number at 0.3 V vs. RHE. The selectivity of H2O2 of NiSe2 is 68% at the same potential. It can also be seen that the selectivity of NC is the lowest, only 38~59%. These results indicate that the combination of NiSe2 and NC is more beneficial to the 2e pathway than that of the pure component. Among the catalysts with different degrees of selenization, NiSe2@NC-H shows the lowest ORR activity (Figure S4), which may be related to the existence form of Ni0.85Se that is not conducive to the reaction under acidic media. Moreover, NiSe2@NC also exhibits the best 2e ORR activity and selectivity.
In addition, the ORR performance of NiSe2@NC under alkaline media (0.1 M KOH) was also evaluated, as shown in Figure S5. NiSe2@NC exhibits an excellent ORR activity with an initial potential of 0.79 V vs. RHE and a disk current density of 2.8 mA cm−2 at 0 V vs. RHE. However, the ring current density is drastically reduced with only ~35% H2O2 selectivity, which is much lower than the performance tested under acidic media. The results suggest that NiSe2@NC is more favorable for electrocatalytic oxygen reduction for H2O2 production under acidic media.
To reveal the effect of the combination of NiSe2 and NC on the improved catalytic performance, other electrochemical properties were further tested. Figure 6a shows the Nyquist plots and fitted charge transfer resistance of the three catalysts to reveal the electronic conductivity [49,50]. The metal-free NC has the smallest semicircular diameter, corresponding to the highest disk current in the LSV curves. In addition, compared with the bare NiSe2, NiSe2@NC also exhibits a smaller semicircular diameter and lower charge transfer resistance, implying a higher electronic conductivity than NiSe2. The cyclic voltammetry curves of the double-layer region (Figure S6) were also performed at different scan rates, which could be used to measure the electrochemical active surface area (ECSA). As seen in Figure 6b, the double-layer capacitance (Cdl) of NC is 2.9 mF cm−2. It is followed by NiSe2@NC with the Cdl value of 0.86 mF cm−2, which is 2.7 times higher than that of NiSe2 (0.32 mF cm−2), revealing a higher active surface area [15]. These results indicate that the combination with NC can not only improve the conductivity of NiSe2, but also expose more active sites.
The Faraday efficiency (FE) and yield of H2O2 are also important indicators to evaluate the practical application of catalysts [51]. Therefore, the actual H2O2 production capacity of NiSe2@NC was evaluated by electrolysis of 7200 s in an H-type cell, combined with UV–Vis spectra of Ce4+ at 319 nm. The electrolytic tests were performed at different potentials to determine the optimal potential of NiSe2@NC. According to Figure S7a,b, the chronoamperometry curves show significantly increased current densities with the negative shift of potential. And the absorbance of Ce4+ also decreases correspondingly, indicating that more H2O2 is produced. This could be confirmed from the calculated Faraday efficiency (FE) and H2O2 concentration of NiSe2@NC in Figure 7a. The highest Faraday efficiency of 81% can be achieved at 0.3 V vs. RHE. Meanwhile, the H2O2 concentration increases continuously as the potential becomes negative, with the maximum of 68 ppm at 0.2 V vs. RHE. In general, NiSe2@NC has the best catalytic performance at 0.3 V vs. RHE. Subsequently, the electrochemical performance of H2O2 production from NC, NiSe2@NC, and NiSe2 was compared at 0.3 V vs. RHE, as shown in Figure 7b. Note that the NiSe2@NC has a higher Faraday efficiency (81%) and H2O2 concentration (53 ppm) than that of NC (59%, 38 ppm) and NiSe2 (73%, 12 ppm), which is consistent with the RRDE trend. It can be seen that NiSe2@NC has the best H2O2 production capacity, which also proves that the combination of NiSe2 and NC plays a great role in enhancing the performance of 2e ORR.
The stability is another important factor to evaluate the performance of catalysts in practical applications. Therefore, the 10 h electrolysis experiments were performed at 0.4 V vs. RHE to evaluate the stability of NiSe2@NC and NiSe2 (Figure 7c). Obviously, the chronoamperometry curve of bare NiSe2 exhibits a much higher loss in current density than NiSe2@NC. According to the XRD patterns in Figure 7d and the wide range of patterns in Figure S8, the crystal structure of NiSe2@NC could not be destroyed after the long-term electrolysis. Combined with the above analysis results, the introduction of NC can significantly improve the stability of the catalytic system in 0.1 M HClO4, which is more conducive to the production of H2O2. Table S1 compares the 2e ORR performance of other reported transition metal chalcogenides. It can be seen that NiSe2@NC is a promising 2e ORR electrocatalyst with a high selectivity and H2O2 production capacity under acidic media.

2.3. Electrocatalytic Mechanism on NiSe2@NC

To further investigate the catalytic mechanism of NiSe2@NC during the ORR process, in situ Raman measurement was used to detect the intermediates produced during the reaction in O2-saturated 0.1 M HClO4. The Raman spectra of different potentials were collected at a 638 nm excitation wavelength. As shown in Figure 8a, the signal peak at 933 cm−1 can be associated to the symmetric stretching mode of ClO4 [52]. Another peak at 463 cm−1 also presents at all potentials, which may be related to the adsorption of oxygen-containing species on the surface of the catalyst, resulting in the formation of a Ni-O bond [53]. Furthermore, only the peaks of the ClO4 and Ni-O bond could be identified at the open circuit potential (OCP). The new peak signal located at 733 cm−1 appears when the potential is shifted negatively to 0.4 V vs. RHE, which is assigned to the vibration of the key intermediate *OOH [54]. And this peak intensity increases with the negative shift of potential, implying the accumulation of *OOH. These results indicate that NiSe2@NC is an effective 2e ORR catalyst that can stabilize *OOH on the surface without further O-O bond breaking.
Based on the above in situ Raman results, density functional theoretical (DFT) calculations were carried out to further illustrate how the introduction of NC enhanced the 2e ORR activity and selectivity. The optimized simulation models of *OOH absorbed on NiSe2 and NiSe2@NC are exhibited in Figures S9 and S10. Undoubtedly, the pathway of the oxygen reduction reaction is mainly determined by the *OOH adsorption strength on the surface of the catalyst, which is also closely related to the electronic state of the active site [33]. Therefore, Bader charge analysis was performed to reveal the different electronic states of NiSe2 and NiSe2@NC. As shown in Figure 8b, the active sites of NiSe2 and NiSe2@NC obtain 0.30 e and 0.35 e, respectively. Therefore, the active site of NiSe2@NC has more electrons and carries more negative charge than NiSe2, which is conducive to the adsorption of O2 and H+ during the 2e ORR reaction pathway. These results verify that the introduction of NC could effectively promote the adsorption and activation of oxygen on the active site to further produce *OOH intermediates.
The ideal 2e ORR reaction requires an optimum balance between the adsorption and desorption of *OOH, with neither too strong nor too weak an adsorption strength. Then, the *OOH adsorption free energies (ΔG*OOH) of the two catalysts were calculated at the equilibrium potential (U = 0.7 V) and standard condition (U = 0 V) to evaluate the adsorption strength of *OOH. As shown in Figure 8c,d, the *OOH adsorption energies of NiSe2@NC at U = 0.7 V and 0 V are 2.70 eV and 4.10 eV, respectively, both of which are closer to the ideal values (3.52 eV and 4.22 eV) than NiSe2 (2.46 eV and 3.86 eV) [55,56]. From the above analysis, the conversion from *OOH to H2O2 production on the surface of the NiSe2@NC catalyst is more favorable than for NiSe2. In addition, the differential charge density distributions between adsorbed *OOH and substrates were also simulated and are displayed in Figure S11. In comparing the two catalysts, the adsorption site of NiSe2 has more positive charges than that of NiSe2@NC, indicating that NiSe2 exhibits stronger charge interactions with *OOH, which may contribute to the breaking of the O-O bond. On the contrary, the *OOH adsorption on NiSe2@NC is moderate, which is conducive to preserving the O-O bond and promoting the next hydrogenation to H2O2.
In conclusion, the above DFT results indicate that the introduction of NC could not only change the electronic state of NiSe2 to promote the adsorption and activation of oxygen, but also optimize the adsorption free energy of *OOH to preserve the O-O bond and convert it to H2O2 during the oxygen reduction reaction. Therefore, NiSe2@NC shows a significantly increased 2e ORR activity and selectivity.

3. Materials and Methods

3.1. Reagent and Chemicals

All reagents were of analytical grade and used without further purification. Zinc nitrate hexahydrate (Zn(NO3)2·6H2O), nickel chloride hexahydrate (NiCl2·6H2O), and potassium hydroxide (KOH) were purchased from Aladdin Bio-technology Co., Ltd., Shanghai, China. Selenium powder (Se), methanol (CH3OH), and anhydrous ethanol (C2H5OH) were obtained from Yuanli Chemical Co., Ltd., Tianjin, China. The compound 2-methylimidazole (2-melm) was purchased from Bailingwei Technology Co., Ltd., Beijing, China. The Nafion solution was purchased from Yingke United Co., Ltd., Tianjin, China.

3.2. Material Synthesis

3.2.1. Synthesis of ZIF-8

The compound ZIF-8 was synthesized based on a previous reported method with a slight adjustment [39]. Typically, 1.487 g of Zn(NO3)2·6H2O and 1.644 g of 2-methylimidazole were dissolved in 50 mL of methanol solution, respectively. After mixing the two solutions, the mixture was stirred rapidly for 30 min and left at room temperature for 24 h. Finally, the white product was centrifuged, washed with ethanol, and fully dried in a vacuum drying oven at 60 °C.

3.2.2. Synthesis of NiSe2@NC

A total of 200 mg of ZIF-8 was ultrasonically dispersed in 60 mL of ethanol. Then, 100 mg of NiCl2·6H2O was added to the above solution and stirred at room temperature for 12 h. The precipitate was centrifuged and dried to obtain Ni-ZIF-8. Subsequently, the Ni-ZIF-8 was transferred to a tube furnace and annealed at 1000 °C for 3 h with a heating rate of 5 °C min−1 under an Ar atmosphere to form Ni-N-C. Finally, a 2:1 mass ratio of Se powder and Ni-N-C were loaded into two porcelain boats. One of the porcelain boats containing Se powder was placed upstream of the tube furnace and annealed at 500 °C for 2 h with a heating rate of 2 °C min−1 under an Ar atmosphere. The samples after pyrolysis were recorded as NiSe2@NC. In addition, the content of NiCl2·6H2O was varied during catalyst preparation by adding 50 mg and 150 mg of NiCl2·6H2O, respectively. Other steps were carried out under the same conditions. The final samples were respectively labeled as NiSe2@NC-L and NiSe2@NC-H, respectively.

3.2.3. Synthesis of NC and NiSe2

The NC catalyst was obtained by the direct calcination of ZIF-8 under the same conditions. The NiSe2 catalyst was synthesized by a one-step hydrothermal reaction. In total, 0.24 g of NiCl2·6H2O, 0.18 g of Se powder, and 3 g of KOH were added to 20 mL of deionized water and stirred for 30 min at room temperature. Subsequently, the mixture was transferred to an oven for 12 h at 150 °C. After the solution was cooled to room temperature, it was washed with deionized water and finally dried in a vacuum drying oven at 60 °C. The resulting samples were denoted as NiSe2.

3.3. Materials Characterization

The morphology and structure of the catalyst were measured by scanning electron microscopy (SEM, S-4800, Hitachi, Tokyo, Japan) and transmission electron microscopy (TEM, JEM-2100F, Japan Electronics, Islamabad, Pakistan). The distribution of elements was determined by energy dispersive X-ray spectroscopy (EDX) using scanning electron microscopy. The crystal structure was measured using X-ray diffraction (XRD, D8 Focus, Bruck, Billerica, MA, USA) using a Cu-Kα radiation source at a scan rate of 5 °C min−1. X-ray photoelectron spectroscopy (XPS, K-Alpha+, Thermo, Waltham, MA, USA) was performed to determine the composition and chemical state. Raman spectra (Horiba, Kyoto, Japan) were measured at 638 nm to determine the phase composition. The optical properties were detected by an ultraviolet–visible spectrophotometer (UV–Vis, T2600, Yoke, Shanghai, China). The crystallites’ size from the XRD pattern could be calculated by the Scherrer equation.

3.4. Electrochemical Measurements

All electrochemical tests in this paper were performed using a CHI 760E electrochemical workstation (Chenhua Instrument Co., Ltd., Shanghai, China) and 0.1 M HClO4. The 2e ORR performance of catalysts was evaluated using a rotating ring disk electrode (RRDE, disk area: 0.247 cm2, platinum ring area: 0.186 cm2) as the working electrode. Carbon rods were used as the counter electrodes and saturated calomel electrodes (SCE) as the reference electrodes to form the test system. The ink was prepared by mixing 2.5 mg of catalyst, 500 μL of ethanol, 500 μL of H2O, and 20 μL of Nafion solution. After ultrasonic dispersion, 10 μL of ink was added to the RRDE surface with the catalyst loading of 0.1 mg cm−2. Before the catalytic performance measurement, cyclic voltammetry (CV) was performed on catalyst-loaded RRDE under a N2 atmosphere, scanning in the potential range from 0 to 0.8 V vs. a reversible hydrogen electrode (RHE) at 50 mV s−1 for around 20 cycles. After the CV curves remained stable, the linear voltametric sweep (LSV) test was performed in O2-saturated 0.1 M HClO4. In the same potential range, LSV curves were measured at a scan rate of 5 mV s−1 and a rotating speed of 1600 rpm. Moreover, the Pt ring voltage was kept at 1.3 V vs. RHE during the test. The selectivity of H2O2 and electron transfer number (n) were calculated according to the following formula:
H 2 O 2 % = 200   ×   I r / [ ( N   ×   I d ) + I r ]
n = 4   ×   I d / ( I d + I r / N )
where Id is the disk current, Ir is the ring current, and N is the collection efficiency of the Pt ring, which could be identified by the LSV curves at different speeds. As shown in Figure S12, the collection efficiency is 37%.
The Faraday efficiency (FE) and yield of H2O2 are determined by the chromogenic reaction of Ce(SO4)2 (2Ce4+ + H2O2 → 2Ce3+ + 2H+ + O2). In the presence of H2O2, Ce(SO4)2 (yellow) can be converted to Ce3+ (colorless), which could reduce the absorbance of UV–Vis at 319 nm. Therefore, the H2O2 production capacity of the catalyst can be measured by the change in absorbance. The standard solution of Ce(SO4)2 with different concentrations was configured by dissolving Ce(SO4)2 in a 0.5 M H2SO4 solution. According to the UV–Vis peak absorbance and Ce4+ concentration, the standard curve is drawn in Figure S13 (y = 4.084x − 0.0312). The electrolytic process is carried out in an H-type cell, as shown in Figure S14. A 1 × 1 cm2 carbon paper supported with the catalyst was used as the working electrode, saturated calomel electrodes as the reference electrodes, and the Pt sheet as the counter electrode. After two hours of electrolysis, 50 μL of electrolytic solution was added to 3 mL of 0.5 mM Ce(SO4)2 solution, and then the change in the peak absorbance of Ce4+ was recorded. The FE and yield of H2O2 can be calculated according to the following formula:
FE % = 200   ×   C   ×   V   ×   F / Q
[ H 2 O 2 ]   ( ppm ) = 1 2   ×   V 1   ×   C before     ( V 1 + V 2 )   ×   C after V 2   ×   34.01
where C is the concentration of the H2O2 product (mol L−1), V is the volume of the solution in the electrolytic cell (L), F is the Faraday constant (96,485 C mol−1), V1 is the volume of Ce4+ before the reaction (mL), Cbefore is the concentration of Ce4+ before the reaction (mmol L−1), V2 is the volume of the added electrolyte solution (mL), and Cafter is the concentration of Ce4+ after the reaction (mmol L−1).
And in this work, all of the measured SCE potentials were converted to the reversible hydrogen electrode (RHE) potentials according to the following formula:
E RHE = E SCE + 0.059   pH + 0.233

3.5. Computational Method

All calculations in this work were performed using the VASP software package (5.4.4). The GGA-PBE function is used to describe the exchange correlation interaction, the PAW method is used to describe the direct interaction between nuclei and valence electrons, and the DFT-D3 method is used to modify the van der Waals interaction. The convergence criteria for energy and force are 10−4 eV and 0.05 eV/Å. The plane wave truncation energy is set to 400 eV, and a 3 × 3 × 1 K-point grid is used to optimize the geometric structure and calculate the frequency and electronic characteristics of the catalyst. The optimal NiSe2 exposed crystal surface is (210), the catalyst is set to a thickness of 3 layers, and the NiSe2 is p (2 × 1) period units. The graphene carrier is a 4 × 3 supercell containing 48 carbon atoms, and a 15 Å vacuum layer is added to reduce the interaction between adjacent layers. The free energy is calculated according to the following formula:
G = E elect + E ZPE TS + G U
where Eelect is the electronic energy calculated by DFT, EZPE and S are the zero-point energy and entropy, T is 298.15 K, and GU = −neU.

4. Conclusions

In summary, we present a nitrogen-doped carbon-coated NiSe2 (NiSe2@NC) electrocatalyst for boosting the H2O2 production from 2e ORR in acidic media. NiSe2@NC exhibited the better 2e ORR activity and selectivity, with the yield of H2O2 being 4.4 times that of bare NiSe2. In addition, the long-term stability of 10 h could be achieved in the electrolysis process. This significantly improved 2e ORR performance was mainly attributed to the synergistic effect of NiSe2 and NC. The introduction of NC could reduce the charge transfer resistance and stabilize the NiSe2 nanoparticles to expose a greater electrochemically active surface area. Moreover, in situ Raman tests and DFT calculations were further performed to reveal the role of the composition of NiSe2 with NC. The introduction of NC could not only adjust the electronic state of NiSe2 to carry more negative charges on the active site, but it could also optimize the adsorption free energy of *OOH, thus further promoting 2e ORR for the production of H2O2. This work provides an effective strategy to modify the transition metal selenides for selectively producing H2O2 by the 2e ORR reaction under acidic media.

Supplementary Materials

The following supporting information can be downloaded at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/catal14060364/s1. Figure S1: (a) SEM image and (b) XRD pattern of ZIF-8; Figure S2: SEM images of (a) NiSe2@NC-L and (b) NiSe2@NC-H. (c) XRD patterns of NiSe2@NC-L and NiSe2@NC-H; Figure S3. XPS full spectrum of NiSe2@NC; Figure S4: (a) RRDE polarization curves, (b) H2O2 selectivity and electron transfer number of NiSe2@NC-L, NiSe2@NC, and NiSe2@NC-H at 1600 rpm in O2-saturated 0.1 M HClO4; Figure S5: (a) RRDE polarization curves and (b) H2O2 selectivity of NiSe2@NC at 1600 rpm in O2-saturated 0.1 M KOH; Figure S6: Cyclic voltammograms of (a) NC, (b) NiSe2@NC, and (c) NiSe2 at different scan rates; Figure S7: (a) Chronoamperometry curves for NiSe2@NC at different potentials for 7200 s in O2-saturated 0.1 M HClO4 and (b) the corresponding UV–Vis spectra; Figure S8: XRD patterns of NiSe2@NC before and after electrolysis over a wide range; Figure S9: The DFT calculation models of NiSe2-OOH, where the yellow, blue, red, and white spheres represent the Se, Ni, O, and H atoms, respectively; Figure S10: The DFT calculation models of NiSe2@NC-OOH, where the yellow, blue, red, white, and gray spheres represent the Se, Ni, O, H, and C atoms, respectively; Figure S11: Differential charge density distributions between adsorbed *OOH and (a) NiSe2 and (b) NiSe2@NC substrates, where the green, gray, and brown spheres represent the Se and Ni atom, respectively, while the cyan and yellow color isosurfaces mean the negative and positive charge, respectively; Figure S12: (a) LSV curves of RRDE at different speeds and (b) linear fitting curve of the disk current and ring current; Figure S13: (a) UV–Vis spectra of different standard Ce4+ solution and (b) the corresponding standard curve. Figure S14: Diagram of an H-type electrolytic cell reaction device; Table S1: Comparison of 2e ORR properties with other transition metal chalcogenides. References [26,27,33,36,37,57,58,59,60] are cited in the Supplementary Materials.

Author Contributions

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

Funding

This research was funded by the National Natural Science Foundation of China (Grant No. 22178266).

Data Availability Statement

Data are available within the article.

Acknowledgments

We are grateful to the analysis and test center of Tianjin University for providing XRD, SEM, and XPS characterizations. We are also grateful to the Donghai Mei group in Tiangong University for providing in situ Raman spectroscopy tests.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Tian, Y.; Deng, D.; Xu, L.; Li, M.; Chen, H.; Wu, Z.; Zhang, S. Strategies for Sustainable Production of Hydrogen Peroxide via Oxygen Reduction Reaction: From Catalyst Design to Device Setup. Nano-Micro Lett. 2023, 15, 122. [Google Scholar] [CrossRef] [PubMed]
  2. Peng, W.; Tan, H.; Liu, X.; Hou, F.; Liang, J. Perspectives on Carbon-Based Catalysts for the Two-Electron Oxygen Reduction Reaction for Electrochemical Synthesis of Hydrogen Peroxide: A Minireview. Energy Fuels 2023, 37, 17863–17874. [Google Scholar] [CrossRef]
  3. Jung, E.; Shin, H.; Hooch Antink, W.; Sung, Y.-E.; Hyeon, T. Recent Advances in Electrochemical Oxygen Reduction to H2O2: Catalyst and Cell Design. ACS Energy Lett. 2020, 5, 1881–1892. [Google Scholar] [CrossRef]
  4. Wang, N.; Ma, S.; Zuo, P.; Duan, J.; Hou, B. Recent Progress of Electrochemical Production of Hydrogen Peroxide by Two-Electron Oxygen Reduction Reaction. Adv. Sci. 2021, 8, 2100076. [Google Scholar] [CrossRef] [PubMed]
  5. Wu, Y.; Sun, J.; Dou, S.; Sun, J. Non-precious metal electrocatalysts for two-electron oxygen electrochemistry: Mechanisms, progress, and outlooks. J. Energy Chem. 2022, 69, 54–69. [Google Scholar] [CrossRef]
  6. Ali, I.; Van Eyck, K.; De Laet, S.; Dewil, R. Recent advances in carbonaceous catalyst design for the in situ production of H2O2 via two-electron oxygen reduction. Chemosphere 2022, 308, 136127. [Google Scholar] [CrossRef] [PubMed]
  7. Zhang, Q.; Zheng, L.; Gu, F.; Wu, J.; Gao, J.; Zhang, Y.-C.; Zhu, X.-D. Recent advances in single-atom catalysts for acidic electrochemical oxygen reduction to hydrogen peroxide. Nano Energy 2023, 116, 108798. [Google Scholar] [CrossRef]
  8. Liu, K.; Li, F.; Zhan, H.; Zhan, S. Recent progress in two-dimensional materials for generation of hydrogen peroxide by two-electron oxygen reduction reaction. Mater. Today Energy 2024, 40, 101500. [Google Scholar] [CrossRef]
  9. Du, M.; Li, D.; Liu, S.; Yan, J. Practical Classification of Catalysts for Oxygen Reduction Reactions: Optimization Strategies and Mechanistic Analysis. Adv. Funct. Mater. 2023, 33, 2301527. [Google Scholar] [CrossRef]
  10. Bhuvanendran, N.; Ravichandran, S.; Xu, Q.; Maiyalagan, T.; Su, H. A quick guide to the assessment of key electrochemical performance indicators for the oxygen reduction reaction: A comprehensive review. Int. J. Hydrog. Energy 2022, 47, 7113–7138. [Google Scholar] [CrossRef]
  11. Xu, X.; Pan, Y.; Zhong, Y.; Ran, R.; Shao, Z. Ruddlesden–Popper perovskites in electrocatalysis. Mater. Horiz. 2020, 7, 2519–2565. [Google Scholar] [CrossRef]
  12. Wang, Z.; Li, Q.-K.; Zhang, C.; Cheng, Z.; Chen, W.; McHugh, E.A.; Carter, R.A.; Yakobson, B.I.; Tour, J.M. Hydrogen Peroxide Generation with 100% Faradaic Efficiency on Metal-Free Carbon Black. ACS Catal. 2021, 11, 2454–2459. [Google Scholar] [CrossRef]
  13. Liu, M.; Su, H.; Cheng, W.; Yu, F.; Li, Y.; Zhou, W.; Zhang, H.; Sun, X.; Zhang, X.; Wei, S.; et al. Synergetic Dual-Ion Centers Boosting Metal Organic Framework Alloy Catalysts toward Efficient Two Electron Oxygen Reduction. Small 2022, 18, 2202248. [Google Scholar] [CrossRef] [PubMed]
  14. Fu, H.; Zhang, N.; Lai, F.; Zhang, L.; Wu, Z.; Li, H.; Zhu, H.; Liu, T. Lattice Strained B-Doped Ni Nanoparticles for Efficient Electrochemical H2O2 Synthesis. Small 2022, 18, 2203510. [Google Scholar] [CrossRef] [PubMed]
  15. Zhang, Y.; Chen, Q.; Guo, A.; Wang, X.; Wang, Y.; Long, Y.; Fan, G. Carbon-nanosheet-driven spontaneous deposition of Au nanoparticles for efficient electrochemical utilizations toward H2O2 generation and detection. Chem. Eng. J. 2022, 445, 136586. [Google Scholar] [CrossRef]
  16. Zhang, J.-Y.; Xia, C.; Wang, H.-F.; Tang, C. Recent advances in electrocatalytic oxygen reduction for on-site hydrogen peroxide synthesis in acidic media. J. Energy Chem. 2022, 67, 432–450. [Google Scholar] [CrossRef]
  17. Zhang, C.; Shen, W.; Guo, K.; Xiong, M.; Zhang, J.; Lu, X. A Pentagonal Defect-Rich Metal-Free Carbon Electrocatalyst for Boosting Acidic O2 Reduction to H2O2 Production. J. Am. Chem. Soc. 2023, 145, 11589–11598. [Google Scholar] [CrossRef] [PubMed]
  18. Huang, X.; Zhang, J.; Luo, G.; Wang, D. Cobalt atoms anchored on nitrogen-doped hollow carbon spheres for efficient electrocatalysis of oxygen reduction to H2O2. J. Phys. Energy 2023, 5, 025001. [Google Scholar] [CrossRef]
  19. Cui, X.; Zhong, L.; Zhao, X.; Xie, J.; He, D.; Yang, X.; Lin, K.; Wang, H.; Niu, L. Ultrafine Co nanoparticles confined in nitrogen-doped carbon toward two-electron oxygen reduction reaction for H2O2 electrosynthesis in acidic media. Chin. Chem. Lett. 2023, 34, 108291. [Google Scholar] [CrossRef]
  20. Sui, W.; Li, W.; Zhang, Z.; Wu, W.; Xu, Z.; Wang, Y. Efficient and durable electrochemical oxygen reduction to H2O2 in acidic media assisted through catalyst layer design. J. Power Sources 2023, 556, 232438. [Google Scholar] [CrossRef]
  21. Wang, X.; Liu, Y.; Liu, Z.; Li, Z.; Zhang, T.; Cheng, Y.; Lei, L.; Yang, B.; Hou, Y. Highly efficient electrosynthesis of H2O2 in acidic electrolyte on metal-free heteroatoms co-doped carbon nanosheets and simultaneously promoting Fenton process. Chin. Chem. Lett. 2024, 35, 108926. [Google Scholar] [CrossRef]
  22. Deng, Z.; Mostaghimi, A.H.B.; Gong, M.; Chen, N.; Siahrostami, S.; Wang, X. Pd 4d Orbital Overlapping Modulation on Au@Pd Nanowires for Efficient H2O2 Production. J. Am. Chem. Soc. 2024, 146, 2816–2823. [Google Scholar] [CrossRef] [PubMed]
  23. Song, M.; Chen, M.; Zhang, C.; Zhang, J.; Liu, W.; Huang, X.; Li, J.; Feng, G.; Wang, D. Modulating the Oxygen Reduction Selectivity in Pt or Pd Chalcogenides via the Ensemble Effect and Electronic Effect. ACS Appl. Mater. Interfaces 2023, 15, 31375–31383. [Google Scholar] [CrossRef]
  24. Li, H.; Wen, P.; Itanze, D.S.; Hood, Z.D.; Adhikari, S.; Lu, C.; Ma, X.; Dun, C.; Jiang, L.; Carroll, D.L.; et al. Scalable neutral H2O2 electrosynthesis by platinum diphosphide nanocrystals by regulating oxygen reduction reaction pathways. Nat. Commun. 2020, 11, 3928. [Google Scholar] [CrossRef] [PubMed]
  25. Sun, X.; Zhu, X.; Wang, Y.; Li, Y. 1T′-MoTe2 monolayer: A promising two-dimensional catalyst for the electrochemical production of hydrogen peroxide. Chin. J. Catal. 2022, 43, 1520–1526. [Google Scholar] [CrossRef]
  26. Sheng, H.; Hermes, E.D.; Yang, X.; Ying, D.; Janes, A.N.; Li, W.; Schmidt, J.R.; Jin, S. Electrocatalytic Production of H2O2 by Selective Oxygen Reduction Using Earth-Abundant Cobalt Pyrite (CoS2). ACS Catal. 2019, 9, 8433–8442. [Google Scholar] [CrossRef]
  27. Yan, L.; Cheng, X.; Wang, Y.; Wang, Z.; Zheng, L.; Yan, Y.; Lu, Y.; Sun, S.; Qiu, W.; Chen, G. Exsolved Co3O4 with tunable oxygen vacancies for electrocatalytic H2O2 production. Mater. Today Energy 2022, 24, 100931. [Google Scholar] [CrossRef]
  28. Han, C.-D.; Zhang, Y.-C.; Zhang, Q.; Wu, J.-T.; Gao, J.; Zou, J.-J.; Zhu, X.-D. NaBH4-induced phase transition of CoSe2 with abundant Se deficiency for acidic oxygen reduction to hydrogen peroxide. Rare Metals 2023, 43, 500–510. [Google Scholar] [CrossRef]
  29. Sheng, H.; Janes, A.N.; Ross, R.D.; Kaiman, D.; Huang, J.; Song, B.; Schmidt, J.R.; Jin, S. Stable and selective electrosynthesis of hydrogen peroxide and the electro-Fenton process on CoSe2 polymorph catalysts. Energy Environ. Sci. 2020, 13, 4189–4203. [Google Scholar] [CrossRef]
  30. Wang, Y.N.; Han, C.D.; Ma, L.; Duan, T.G.; Du, Y.; Wu, J.T.; Zou, J.J.; Gao, J.; Zhu, X.D.; Zhang, Y.C. Recent Progress of Transition Metal Selenides for Electrochemical Oxygen Reduction to Hydrogen Peroxide: From Catalyst Design to Electrolyzers Application. Small 2024, 21, 2309448. [Google Scholar] [CrossRef]
  31. Cao, W.; Shen, Q.; Men, D.; Ouyang, B.; Sun, Y.; Xu, K. Phase engineering of iron group transition metal selenides for water splitting. Mat. Chem. Front. 2023, 7, 4865–4879. [Google Scholar] [CrossRef]
  32. Gong, Y.; Li, Y.; Li, Y.; Liu, M.; Bai, Y.; Wu, C. Metal Selenides Anode Materials for Sodium Ion Batteries: Synthesis, Modification, and Application. Small 2022, 19, 2206194. [Google Scholar] [CrossRef] [PubMed]
  33. Sun, Q.; Xu, G.; Xiong, B.; Chen, L.; Shi, J. Anion-tuned nickel chalcogenides electrocatalysts for efficient 2e ORR towards H2O2 production in acidic media. Nano Res. 2022, 16, 4729–4735. [Google Scholar] [CrossRef] [PubMed]
  34. Yu, L.; Pang, X.; Tian, Z.; Wang, S.; Feng, L. Fe-doped NiSe2 nanorods for enhanced urea electrolysis of hydrogen generation. Electrochim. Acta 2023, 440, 141724. [Google Scholar] [CrossRef]
  35. Fan, S.; Li, G.; Yang, G.; Guo, X.; Niu, X. NiSe2 nanooctahedra as anodes for high-performance sodium-ion batteries. New J. Chem. 2019, 43, 12858–12864. [Google Scholar] [CrossRef]
  36. Zhang, L.; Liang, J.; Yue, L.; Dong, K.; Xu, Z.; Li, T.; Liu, Q.; Luo, Y.; Liu, Y.; Gao, S.; et al. CoTe nanoparticle-embedded N-doped hollow carbon polyhedron: An efficient catalyst for H2O2 electrosynthesis in acidic media. J. Mater. Chem. A 2021, 9, 21703–21707. [Google Scholar] [CrossRef]
  37. Zhang, L.; Liang, J.; Yue, L.; Xu, Z.; Dong, K.; Liu, Q.; Luo, Y.; Li, T.; Cheng, X.; Cui, G.; et al. N-doped carbon nanotubes supported CoSe2 nanoparticles: A highly efficient and stable catalyst for H2O2 electrosynthesis in acidic media. Nano Res. 2021, 15, 304–309. [Google Scholar] [CrossRef] [PubMed]
  38. Xu, X.; Sun, H.; Jiang, S.P.; Shao, Z. Modulating metal–organic frameworks for catalyzing acidic oxygen evolution for proton exchange membrane water electrolysis. SusMat 2021, 1, 460–481. [Google Scholar] [CrossRef]
  39. Guo, R.; Shi, W.; Liu, W.; Yang, X.; Xie, Y.; Yang, T.; Xiao, J. Ultralow noble metals doping enables metal-organic framework derived Ni(OH)2 nanocages as efficient water oxidation electrocatalysts. Chem. Eng. J. 2022, 429, 132478. [Google Scholar] [CrossRef]
  40. Qu, J.; Bai, Y.; Li, X.; Song, K.; Zhang, S.; Wang, X.; Wang, X.; Dai, S. Rational design of NiSe2@rGO nanocomposites for advanced hybrid supercapacitors. J. Mater. Res. Technol. JMRT 2021, 15, 6155–6161. [Google Scholar] [CrossRef]
  41. Bai, J.; Ge, W.; Zhou, P.; Xu, P.; Wang, L.; Zhang, J.; Jiang, X.; Li, X.; Zhou, Q.; Deng, Y. Precise constructed atomically dispersed Fe/Ni sites on porous nitrogen-doped carbon for oxygen reduction. J. Colloid Interface Sci. 2022, 616, 433–439. [Google Scholar] [CrossRef] [PubMed]
  42. Chen, W.-C.; Yang, G.; Zhao, Y.; Yuan, G.-Q.; Ye, J.-S.; Liu, H.-Y.; Xiao, X.-Y. Porous carbon polyhedrons with exclusive Metal-NX moieties for efficient oxygen reduction reaction. Int. J. Hydrogen Energy 2021, 46, 39882–39891. [Google Scholar] [CrossRef]
  43. Wu, Y.; Ge, L.; Veksha, A.; Lisak, G. Cobalt and nitrogen co-doped porous carbon/carbon nanotube hybrids anchored with nickel nanoparticles as high-performance electrocatalysts for oxygen reduction reactions. Nanoscale 2020, 12, 13028–13033. [Google Scholar] [CrossRef] [PubMed]
  44. Long, Y.; Li, Q.; Zhang, Z.; Zeng, Q.; Liu, D.; Zhao, L.; Liu, Y.; Li, Y.; Zhang, Y.; Ji, K.; et al. Coupling MoSe2 with Non-Stoichiometry Ni0.85Se in Carbon Hollow Nanoflowers for Efficient Electrocatalytic Synergistic Effect on Li-O2 Batteries. Small 2023, 20, 2304882. [Google Scholar] [CrossRef] [PubMed]
  45. Zhang, Y.; Yun, S.; Sun, M.; Wang, X.; Zhang, L.; Dang, J.; Yang, C.; Yang, J.; Dang, C.; Yuan, S. Implanted metal-nitrogen active sites enhance the electrocatalytic activity of zeolitic imidazolate zinc framework-derived porous carbon for the hydrogen evolution reaction in acidic and alkaline media. J. Colloid Interface Sci. 2021, 604, 441–457. [Google Scholar] [CrossRef] [PubMed]
  46. Wang, H.; Wei, L.; Shen, J. Metal-free catalyst for efficient pH-universal oxygen reduction electrocatalysis in microbial fuel cell. J. Electroanal. Chem. 2022, 911, 116233. [Google Scholar] [CrossRef]
  47. Liang, Z.; Tu, H.; Zhang, K.; Kong, Z.; Huang, M.; Xu, D.; Liu, S.; Wu, Y.; Hao, X. Self-supporting NiSe2@BCNNTs electrode for High-Performance sodium ion batteries. Chem. Eng. J. 2022, 437, 135421. [Google Scholar] [CrossRef]
  48. Xiao, X.; Ni, L.; Chen, G.; Ai, G.; Li, J.; Qiu, T.; Liu, X. Two-dimensional NiSe2 nanosheets on carbon fiber cloth for high-performance lithium-ion batteries. J. Alloy. Compd. 2020, 821, 153218. [Google Scholar] [CrossRef]
  49. Ding, L.; Zhao, J.; Bao, Z.; Zhang, S.; Shi, H.; Liu, J.; Wang, G.; Peng, X.; Zhong, X.; Wang, J. Synchronous generation of green oxidants H2O2 and O3 by using a heterojunction bifunctional ZnO/ZnS@C electrocatalyst. J. Mater. Chem. A 2023, 11, 3454–3463. [Google Scholar] [CrossRef]
  50. Mohamed, I.M.A.; Kanagaraj, P.; Yasin, A.S.; Iqbal, W.; Liu, C. Electrochemical impedance investigation of urea oxidation in alkaline media based on electrospun nanofibers towards the technology of direct-urea fuel cells. J. Alloy. Compd. 2020, 816, 152513. [Google Scholar] [CrossRef]
  51. Hu, Y.; Zhang, J.; Shen, T.; Li, Z.; Chen, K.; Lu, Y.; Zhang, J.; Wang, D. Efficient Electrochemical Production of H2O2 on Hollow N-Doped Carbon Nanospheres with Abundant Micropores. ACS Appl. Mater. Interfaces 2021, 13, 29551–29557. [Google Scholar] [CrossRef] [PubMed]
  52. Sun, Y.L.; Yue, M.F.; Chen, H.Q.; Ze, H.J.; Wang, Y.H.; Dong, J.C.; Tian, Z.Q.; Fang, P.P.; Li, J.F. Exploring the Effect of Pd on the Oxygen Reduction Performance of Pt by In Situ Raman Spectroscopy. Anal. Chem. 2022, 94, 4779–4786. [Google Scholar] [CrossRef] [PubMed]
  53. Wang, Y.; Yang, H.; Liu, Z.; Yin, K.; Wang, Z.; Huang, H.; Liu, Y.; Kang, Z.; Chen, Z. Efficient hydrogen peroxide production enabled by exploring layered metal telluride as two electron oxygen reduction reaction electrocatalyst. J. Energy Chem. 2023, 87, 247–255. [Google Scholar] [CrossRef]
  54. Jiang, C.; Fei, Y.-F.; Xu, W.; Bao, Z.; Shao, Y.; Zhang, S.; Hu, Z.-T.; Wang, J. Synergistic effects of Bi2O3 and Ta2O5 for efficient electrochemical production of H2O2. Appl. Catal. B Environ. 2023, 334, 122867. [Google Scholar] [CrossRef]
  55. Wang, Y.M.; Huang, H.; Wu, J.; Yang, H.Y.; Kang, Z.H.; Liu, Y.; Wang, Z.W.; Menezes, P.W.; Chen, Z.L. Charge-Polarized Selenium Vacancy in Nickel Diselenide Enabling Efficient and Stable Electrocatalytic Conversion of Oxygen to Hydrogen Peroxide. Adv. Sci. 2023, 10, 10. [Google Scholar] [CrossRef] [PubMed]
  56. Wu, J.; Hou, M.; Chen, Z.; Hao, W.; Pan, X.; Yang, H.; Cen, W.; Liu, Y.; Huang, H.; Menezes, P.W.; et al. Composition Engineering of Amorphous Nickel Boride Nanoarchitectures Enabling Highly Efficient Electrosynthesis of Hydrogen Peroxide. Adv. Mater. 2022, 34, 2202995. [Google Scholar] [CrossRef] [PubMed]
  57. Ross, R.D.; Sheng, H.; Parihar, A.; Huang, J.; Jin, S. Compositionally Tuned Trimetallic Thiospinel Catalysts for Enhanced Electrosynthesis of Hydrogen Peroxide and Built-In Hydroxyl Radical Generation. ACS Catal. 2021, 11, 12643–12650. [Google Scholar] [CrossRef]
  58. Zhang, Y.; Jiang, H.; Zhang, C.; Feng, Y.; Feng, H.; Zhu, S.; Hu, J. High-efficiency oxygen reduction by late transition metal oxides to produce H2O2. J. Mater. Chem. A 2024, 12, 6123–6133. [Google Scholar] [CrossRef]
  59. Xie, J.; Zhong, L.; Yang, X.; He, D.; Lin, K.; Chen, X.; Wang, H.; Gan, S.; Niu, L. Phosphorous and selenium tuning Co-based non-precious catalysts for electrosynthesis of H2O2 in acidic media. Chin. Chem. Lett. 2024, 35, 108472. [Google Scholar] [CrossRef]
  60. Zheng, Y.R.; Hu, S.; Zhang, X.L.; Ju, H.; Wang, Z.; Tan, P.J.; Wu, R.; Gao, F.Y.; Zhuang, T.; Zheng, X.; et al. Black Phosphorous Mediates Surface Charge Redistribution of CoSe2 for Electrochemical H2O2 Production in Acidic Electrolytes. Adv. Mater. 2022, 34, 2205414. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of fabrication process of NiSe2@NC catalyst.
Figure 1. Schematic illustration of fabrication process of NiSe2@NC catalyst.
Catalysts 14 00364 g001
Figure 2. SEM images of (a) NC, (b) NiSe2@NC, and (c) NiSe2. TEM images of (d) NC, (e) NiSe2@NC, and (f) NiSe2. (g) HRTEM image of NiSe2@NC. (hk) The SEM elemental mapping of NiSe2@NC.
Figure 2. SEM images of (a) NC, (b) NiSe2@NC, and (c) NiSe2. TEM images of (d) NC, (e) NiSe2@NC, and (f) NiSe2. (g) HRTEM image of NiSe2@NC. (hk) The SEM elemental mapping of NiSe2@NC.
Catalysts 14 00364 g002
Figure 3. (a) XRD patterns and (b) Raman spectra of NC, NiSe2@NC, and NiSe2.
Figure 3. (a) XRD patterns and (b) Raman spectra of NC, NiSe2@NC, and NiSe2.
Catalysts 14 00364 g003
Figure 4. (a) C 1s diagram, (b) N 1s diagram, (c) Ni 2p diagram, and (d) Se 3d diagram of NiSe2@NC, where the circles represent the raw data and the red lines represent the fitted data.
Figure 4. (a) C 1s diagram, (b) N 1s diagram, (c) Ni 2p diagram, and (d) Se 3d diagram of NiSe2@NC, where the circles represent the raw data and the red lines represent the fitted data.
Catalysts 14 00364 g004
Figure 5. (a) RRDE polarization curves, (b) ring current densities at different potentials, (c) H2O2 selectivity, and (d) electron transfer number of NC, NiSe2@NC, and NiSe2 at 1600 rpm in O2-saturated 0.1 M HClO4.
Figure 5. (a) RRDE polarization curves, (b) ring current densities at different potentials, (c) H2O2 selectivity, and (d) electron transfer number of NC, NiSe2@NC, and NiSe2 at 1600 rpm in O2-saturated 0.1 M HClO4.
Catalysts 14 00364 g005
Figure 6. (a) Nyquist plots and (b) double-layer capacitance measured for NC, NiSe2@NC, and NiSe2.
Figure 6. (a) Nyquist plots and (b) double-layer capacitance measured for NC, NiSe2@NC, and NiSe2.
Catalysts 14 00364 g006
Figure 7. (a) Faraday efficiency and H2O2 concentration of NiSe2@NC at different potentials for 7200 s in O2-saturated 0.1 M HClO4. (b) Faraday efficiency and H2O2 concentration of NC, NiSe2@NC, NiSe2 at 0.3 V vs. RHE. (c) Long-term stability of NiSe2@NC and NiSe2 at 0.4 V vs. RHE for 10 h. (d) XRD patterns of NiSe2@NC before and after electrolysis.
Figure 7. (a) Faraday efficiency and H2O2 concentration of NiSe2@NC at different potentials for 7200 s in O2-saturated 0.1 M HClO4. (b) Faraday efficiency and H2O2 concentration of NC, NiSe2@NC, NiSe2 at 0.3 V vs. RHE. (c) Long-term stability of NiSe2@NC and NiSe2 at 0.4 V vs. RHE for 10 h. (d) XRD patterns of NiSe2@NC before and after electrolysis.
Catalysts 14 00364 g007
Figure 8. (a) In situ Raman spectra of NiSe2@NC in O2-saturated 0.1 M HClO4, where the * in *OOH represents the active site of the reaction. (b) Bader charges of NiSe2 and NiSe2@NC. Free energy diagram for 2e ORR on NiSe2 and NiSe2@NC at (c) U = 0.7 V and (d) U = 0 V.
Figure 8. (a) In situ Raman spectra of NiSe2@NC in O2-saturated 0.1 M HClO4, where the * in *OOH represents the active site of the reaction. (b) Bader charges of NiSe2 and NiSe2@NC. Free energy diagram for 2e ORR on NiSe2 and NiSe2@NC at (c) U = 0.7 V and (d) U = 0 V.
Catalysts 14 00364 g008
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

Cheng, Q.; Ding, H.; Chen, L.; Dong, J.; Yu, H.; Yan, S.; Wang, H. Modification of NiSe2 Nanoparticles by ZIF-8-Derived NC for Boosting H2O2 Production from Electrochemical Oxygen Reduction in Acidic Media. Catalysts 2024, 14, 364. https://0-doi-org.brum.beds.ac.uk/10.3390/catal14060364

AMA Style

Cheng Q, Ding H, Chen L, Dong J, Yu H, Yan S, Wang H. Modification of NiSe2 Nanoparticles by ZIF-8-Derived NC for Boosting H2O2 Production from Electrochemical Oxygen Reduction in Acidic Media. Catalysts. 2024; 14(6):364. https://0-doi-org.brum.beds.ac.uk/10.3390/catal14060364

Chicago/Turabian Style

Cheng, Qiaoting, Hu Ding, Lang Chen, Jiatong Dong, Hao Yu, Shen Yan, and Hua Wang. 2024. "Modification of NiSe2 Nanoparticles by ZIF-8-Derived NC for Boosting H2O2 Production from Electrochemical Oxygen Reduction in Acidic Media" Catalysts 14, no. 6: 364. https://0-doi-org.brum.beds.ac.uk/10.3390/catal14060364

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